23.01.2015 Views

Landscapes Forest and Global Change - ESA - Escola Superior ...

Landscapes Forest and Global Change - ESA - Escola Superior ...

Landscapes Forest and Global Change - ESA - Escola Superior ...

SHOW MORE
SHOW LESS
  • No tags were found...

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

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

<strong>L<strong>and</strong>scapes</strong><br />

<strong>Forest</strong><br />

<strong>and</strong> <strong>Global</strong><br />

<strong>Change</strong><br />

New Frontiers<br />

in Management,<br />

Conservation<br />

<strong>and</strong> Restoration<br />

Proceedings<br />

Edited by<br />

João Carlos Azevedo<br />

Manuel Feliciano<br />

José Castro<br />

Maria Alice Pinto<br />

IUFRO L<strong>and</strong>scape Ecology Working Group<br />

International Conference<br />

Bragança · Portugal<br />

September 21 to 27, 2010


<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong><br />

New Frontiers in Management, Conservation <strong>and</strong> Restoration


The contributions to this volume have not been peer-reviewed. Only minor changes in<br />

format may have been carried out by the editors.<br />

Title: <strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong><br />

Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology Working Group International<br />

Conference, September 21-27, 2010, Bragança, Portugal.<br />

Editors: João Carlos Azevedo, Manuel Feliciano, José Castro & Maria Alice Pinto<br />

Published by: Instituto Politécnico de Bragança<br />

Apartado 1038, 5301-854 Bragança, Portugal<br />

http://www.ipb.pt<br />

Printed by: Serviços de Imagem do Instituto Politécnico de Bragança<br />

ISBN: 978-972-745-110-4<br />

Cover design: Atilano Suarez, Serviços de Imagem do Instituto Politécnico de Bragança


<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong><br />

New Frontiers in Management, Conservation <strong>and</strong> Restoration<br />

Proceedings of the<br />

IUFRO L<strong>and</strong>scape Ecology Working Group International Conference<br />

September 21-27, 2010<br />

Bragança, Portugal<br />

Edited by<br />

João Carlos Azevedo<br />

Manuel Feliciano<br />

José Castro<br />

Maria Alice Pinto<br />

Instituto Politécnico de Bragança, Portugal<br />

September, 2010


Table of contents<br />

Foreword<br />

Preface<br />

xiii<br />

xiv<br />

Section 1 1<br />

Scaling in l<strong>and</strong>scape analysis<br />

Environmental drivers of benthic communities: the importance of l<strong>and</strong>scape metrics 2<br />

Rui Cortes, Simone Var<strong>and</strong>as, Samantha Hughes, Marco Magalhães & Amílcar Teixeira<br />

Habitat suitability models for two species of forest raptors in Catalonia. Methodological consequences 8<br />

related with different scales <strong>and</strong> data sources<br />

David Sánchez de Ron, Lluís Brotons, Santiago Saura & José M. García del Barrio<br />

Section 2 14<br />

Patterns <strong>and</strong> processes in changing l<strong>and</strong>scapes<br />

Recent relations between forestry <strong>and</strong> agriculture in Pol<strong>and</strong> - the rural l<strong>and</strong>scape on the axis of 15<br />

openness <strong>and</strong> enclosure<br />

Barbara Bożętka<br />

A physiotope-based model for ecoregions for the the nationwide ecosystem management of Japan 21<br />

Siew Fong Chen, Tadashi Masuzawa, Yukihiro Morimoto & Hajime Ise<br />

Risk areas to flooding in the Hidrographical Basin of Arroio dos Pereiras in Irati, PR, Brazil 27<br />

Vivian Dallagnol de Campos & Sílvia Méri Carvalho<br />

Ecological aspects of soils deflation development in agrol<strong>and</strong>scapes of the south-east of the western Siberian plain 33<br />

N.S. Evseeva & Z.N. Kvasnikova<br />

A regionally adaptable approach of l<strong>and</strong>scape assessment using l<strong>and</strong>scape metrics within the 2D cellular 36<br />

automaton “Pimp your l<strong>and</strong>scape”<br />

Susanne Frank, Christine Fürst, Carsten Lorz, Lars Koschke & Franz Makeschin<br />

The effect of l<strong>and</strong> cover changes (1960’s-2003) on the habitat morphological spatial pattern <strong>and</strong> population 42<br />

viability of Baird’s tapir in Laguna Lachuá National Park Influence Zone, Guatemala<br />

Manolo García, Fern<strong>and</strong>o Castillo, Raquel Leonardo, Liza García & Ivonne Gómez<br />

Spatial pattern of soil macrofauna biodiversity in wildlife refugee of Karkhe in Southwestern Iran 47<br />

Shaieste Gholami, Seied Mohsen Hosseini , Jahangard Mohammadi & Abdolrassoul Salman Mahini<br />

The influence of spatial structure on natural regeneration <strong>and</strong> biodiversity in Mediterranean pine plantations: 52<br />

a nested l<strong>and</strong>scape approach<br />

P. González-Moreno, J.L. Quero, L. Poorter, F.J. Bonet & R. Zamora<br />

Impact of changing cultivation systems on the l<strong>and</strong>scape structure of La Gamba, southern Costa Rica 58<br />

Tamara Höbinger, Stefan Schindler & Anton Weissenhofer<br />

Can lichen functional diversity be a good indicator of macroclimatic conditions 64<br />

Paula Matos, Pedro Pinho, Esteve Llop & Cristina Branquinho<br />

Projections of shifts in species distributions: assessing the influence of macro-climate <strong>and</strong> local processes 70<br />

Eliane S. Meier, Felix Kienast & Niklaus E. Zimmermann<br />

Application of remote sensing to assess the wildfire impact on the natural vegetation recovery <strong>and</strong> 74<br />

l<strong>and</strong>scape structure in the Mediterranean forest of South eastern Spain<br />

H. Moutahir , J.F. Bellot, A.Bonet, M.J. Baeza & I. Touhami<br />

The use of Voronoi tessellation to characterize sapling populations 79<br />

Ciprian Palaghianu


Spatial <strong>and</strong> temporal dynamics <strong>and</strong> future trends of change in the coastal l<strong>and</strong>scape of La Araucania, Chile 85<br />

Fern<strong>and</strong>o Peña-Cortés, Daniel Rozas, Enrique Hauenstein, Carlos Bertrán, Jaime Tapia & Marco Cisternas<br />

Multi-temporal analysis of forest l<strong>and</strong>scape fragmentation in the North East of Madagascar 91<br />

F.M. Rabenilalana, L.G. Rajoelison, J.P. Sorg, J.L. Pfund & H. Rakoto Ratsimba<br />

Patterns of afforestation process in ab<strong>and</strong>oned agriculture l<strong>and</strong> in Latvia 97<br />

Anda Ruskule, Olģerts Nikodemus, Zane Kasparinska & Raimonds Kasparinskis<br />

Nitrogen retranclocation in pure <strong>and</strong> mixed plantations of Populus deltoides <strong>and</strong> Alnus subcordata 103<br />

Ehsan Sayad<br />

Evaluation of the ecosystem service in the forest formations of Biosphere Reserve “Srebarna”, 107<br />

Northeastern Bulgaria<br />

Georgi Zhelezov<br />

Section 3 113<br />

Disturbances in changing l<strong>and</strong>scapes<br />

Human-caused forest fire in Mediterranean ecosystems of Chile: modelling l<strong>and</strong>scape spatial patterns 114<br />

on forest fire occurrence<br />

Adison Altamirano, Christian Salas & Valeska Yaitul<br />

Are changes in fire regime threatening cork oak-shrubl<strong>and</strong> mosaics 119<br />

Thomas Curt & Juli Pausas<br />

<strong>Forest</strong> management <strong>and</strong> climate, through l<strong>and</strong>scape structure, affect the potential for insect outbreak 124<br />

Richard A Fleming, Allan L. Carroll, Jean-Noël C<strong>and</strong>au & Philippe Dreyfus<br />

How important are riparian forests to aquatic foodwebs in agricultural watersheds of north-central Ohio, USA 129<br />

P. Charles Goebel, Charles W. Goss, Virginie Bouchard & Lance R. Williams<br />

Merger of three modeling approaches to assess potential effects of climate change on trees in the eastern 135<br />

United States<br />

Louis R. Iverson, Anantha M. Prasad, Stephen N. Matthews & Matthew P. Peters<br />

Immediate effects of typhoon disturbance <strong>and</strong> artificial thinning on understory light environments 141<br />

in two subtropical forests in Taiwan<br />

Teng-Chiu Lin, Kuo-Chuan Lin, Jeen-Liang Hwong & Hsueh-Fang Wang<br />

Impact of hemlock decline on successional pathways <strong>and</strong> ecosystem function at multiple scales in forests 147<br />

of the central Appalachians, USA<br />

Katherine L. Martin & P. Charles Goebel<br />

Indirect estimation of l<strong>and</strong>scape uses by Lama guanicoe <strong>and</strong> domestic herbivorous through the study of diet 153<br />

composition in South Patagonia<br />

Guillermo Martínez Pastur, Rosina Soler Esteban, María Vanessa Lencinas & Laura Borrelli<br />

Effects of endozoochorous seed dispersal on the soil seed bank <strong>and</strong> vegetation in the woodl<strong>and</strong> area 159<br />

E.W Saragih, Jan Bokdam & Wim Braakhekke<br />

Anthropogenic l<strong>and</strong>scape changes <strong>and</strong> the conservation of woodl<strong>and</strong> caribou in British Columbia, Canada 165<br />

Libby R.P. Williamson, Chris J. Johnson & Dale R. Seip<br />

Modeling feedbacks between avalanches <strong>and</strong> forests under a changing environment in the Swiss Alps 171<br />

Natalie Zurbriggen, Heike Lischke, Peter Bebi & Harald Bugmann<br />

Section 4 177<br />

Biodiversity conservation <strong>and</strong> planning in changing l<strong>and</strong>scapes<br />

Ecotourism <strong>and</strong> Controlling <strong>Forest</strong> St<strong>and</strong> Damages. Case study: Varzegan district North West Iran 178<br />

M. Akbarzadeh, J. Ajalli & E. Kouhgardi


Fine-scale mapping of High Nature Value farml<strong>and</strong>s: novel approaches to improve the management 182<br />

of rural biodiversity <strong>and</strong> ecosystem services<br />

Cláudia Carvalho-Santos, Rob Jongman, Joaquim Alonso & João Honrado<br />

Wood macrolichen Lobaria pulmonaria on chestnut tree crops: the case study of Roccamonfina park 188<br />

(Campania region - Italy)<br />

I. Catalano, A. Mingo, A. Migliozzi, S. Sgambato & G.G. Aprile<br />

Extent <strong>and</strong> characteristics of mire habitats in Galicia (NW Iberian Peninsula): implications for their conservation 194<br />

<strong>and</strong> management<br />

Ramón Alberto Diaz-Varela, Pablo Ramil-Rego, Manuel Antonio Rodríguez-Guitián & Carmen Cillero-Castro<br />

Assessment of conservation status in managed chestnut forest by means of l<strong>and</strong>scape metrics, 200<br />

physiographic parameters <strong>and</strong> textural features<br />

Ramón Alberto Diaz-Varela, Pedro Álvarez-Álvarez, Emilio Diaz-Varela & Silvia Calvo-Iglesias<br />

GIS analysis of the Antidote Programme in Portugal 206<br />

Patrícia A.E. Diogo & José Aranha<br />

Investigation on Species diversity by Inventory in Caspian <strong>Forest</strong>s (Case study: Gorazbon District- Noshahr, Iran) 211<br />

Vahid Etemad & Nazi Avani<br />

The impact of the matrix on species movement: systematic review <strong>and</strong> meta-analysis 215<br />

A.E. Eycott, G.B. Stewart, G. Br<strong>and</strong>t, L. M. Buyung-Ali, D. E. Bowler, K. Watts & A.S. Pullin<br />

Eight years of development of a silvopastoral system: effects on floristic diversity 221<br />

E. Fernández-Núñez, A. Rigueiro-Rodríguez & M.R. Mosquera-Losada<br />

Participatory action research concerning the l<strong>and</strong>scape use by a native cervid in a wetl<strong>and</strong> of the Plata Basin, 227<br />

Argentina<br />

Natalia Fracassi & Daniel Somma<br />

Fortresses <strong>and</strong> fragments: impacts of fragmentation in a forest park l<strong>and</strong>scape 233<br />

Joel Hartter, Sadie J. Ryan, Jane Southworth & Colin A. Chapman<br />

L<strong>and</strong>scape structure of a conservation area <strong>and</strong> its surroundings in Minas Gerais State, Brazil 239<br />

Marise Barreiros Horta, Carla Araújo Simões, Eduardo Christófaro de Andrade & Luciana Eler França<br />

Ecological factors influencing beta diversity at two spatial scales in a tropical dry forest of the Yucatán Peninsula 245<br />

J. Omar López-Martínez, J. Luis Hernández-Stefanoni & Juan Manuel Dupuy<br />

Impact of roads on ungulate species: a preliminary approach in Portugal 251<br />

Sara Marques, Catarina Ferreira, Rogério Rodrigues, Jorge Cancela & Carlos Fonseca<br />

Electrical network hazard assessment for the avifauna in Portugal 257<br />

Miguel G. C. Martins & José Aranha<br />

Modeling l<strong>and</strong> use/cover change <strong>and</strong> biodiversity conservation in Mexico 262<br />

Jean-François Mas, Azucena Pérez Vega, Keith Clarke & Víctor Sánchez-Cordero<br />

Evaluating an old sustainable national forest in south Brazil to decide their conservation status 268<br />

Rosemeri Segecin Moro & Tiaro Katu Pereira<br />

A methodological proposal for restoration of forests in southern Brazil 274<br />

Rosemeri Segecin Moro & Carlos Hugo Rocha<br />

L<strong>and</strong>scape integration of Mediterranean reforestations: identification of best practices in Madrid region 279<br />

Victoria Núñez, María Dolores Velarde & Antonio García-Abril<br />

Conservation of priority bird species in a protected area of central Greece using geographical location <strong>and</strong> GIS 285<br />

Sofia G. Plexida & Athanassios I. Sfougaris<br />

<strong>Change</strong>s in structure <strong>and</strong> composition of forest st<strong>and</strong>s at regional <strong>and</strong> national level in the last four decades - 291<br />

a consequence of environmental, natural or social factors<br />

Ales Poljanec & Andrej Boncina


Tengmalm owl (Aegolius funereus) <strong>and</strong> Pygmy owl (Glaussidium passerinum) as a surrogate for biodiversity 297<br />

value in the French Alps <strong>Forest</strong>s<br />

Mathilde Redon & S<strong>and</strong>ra Luque<br />

Contribution to the characterization of Gentiana pneumonanthe L. <strong>and</strong> Maculinea alcon L. distribution 303<br />

in the Alvão Natural Park<br />

M. da Conceição C. Rodrigues, Paula M.S.O. Arnaldo & José Aranha<br />

Using Business <strong>and</strong> Biodiversity to put Conservation into practice: The Herdade do Esporão Case study 306<br />

M.C. Silva, S. Antunes, N. Oliveira & F. Gouveia<br />

Legal efficiency <strong>and</strong> cumulative effects in environmentally protected area 312<br />

Talita Nogueira Terra & Rozely Ferreira dos Santos<br />

Relationships among l<strong>and</strong>scape structure, climate <strong>and</strong> rare woody species richness in the tropical forests of the 317<br />

Yucatan Peninsula, Mexico<br />

Erika Tetetla-Rangel, J. Luis Hernández-Stefanoni & Juan Manuel Dupuy<br />

Economic estimations in using of the l<strong>and</strong>scapes planning 323<br />

M. Tsibulnikova<br />

Section 5 330<br />

Monitoring l<strong>and</strong>scape change<br />

Mapping <strong>and</strong> Monitoring l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> –use changing using RS <strong>and</strong> GIS. Case study: Kaleybar Iran 331<br />

M. Akbarzadeh & E. Kouhgardi<br />

<strong>Forest</strong> patches in agricultural l<strong>and</strong>scapes (loess areas of SE Pol<strong>and</strong>) 336<br />

Bogusława Baran-Zgłobicka & Wojciech Zgłobicki<br />

Integrating esthetical <strong>and</strong> ecological values at the central Asia l<strong>and</strong>scape change 340<br />

Joana Beldade & Thomas Panagopoulos<br />

The impact of legislation on the dynamics of l<strong>and</strong> use the River Basin Cará-Cará, Ponta Grossa-PR/Brazil, 346<br />

in period from 1980 to 2007<br />

Silvia Méri Carvalho & Andreza Rocha de Freitas<br />

Quantitative assessment of temporal dynamics in altitudinal-driven ecotones in a section of Valtellina Italian Alps 352<br />

Ramón Alberto Diaz-Varela, Silvia Calvo-Iglesias, Michele Meroni & Roberto Colombo<br />

L<strong>and</strong> use changes in Portugal between 1990 <strong>and</strong> 2005 358<br />

Inês Duarte & Francisco Castro Rego<br />

Simulation of climate scenarios for the region of Campos Gerais, State of Parana, Brazil 364<br />

Jorim Sousa das Virgens Filho & Maysa de Lima Leite<br />

L<strong>and</strong> use changes <strong>and</strong> mixed forest dynamics. The case of Montiferru Mountains (Sardinia, Italy) (1955-2006) 370<br />

Francisco Javier Gómez, Martí Boada & Diego Varga<br />

Analysis of rainfall in the Vila Velha State Park, State of Parana, Southern Brazil, 376<br />

in the period between 1954 <strong>and</strong> 2001<br />

Maysa de Lima Leite & Jorim Sousa das Virgens Filho<br />

Identification of climatic trends for some localities in the Southern Region of Campos Gerais <strong>and</strong> surroundings, 381<br />

State of Parana, Brazil, through the analysis of historical data of rainfall <strong>and</strong> temperature<br />

Maysa de Lima Leite & Jorim Sousa das Virgens Filho<br />

Monitoring vulnerability of the Spanish forest l<strong>and</strong>scapes: the SISPARES approach 386<br />

Marta Ortega, Valentín Gomez, Jose Manuel García del Barrio & Ramon Elena-Rosselló<br />

L<strong>and</strong>scape transformations seen through the historical cartography: Sardinia as case study 392<br />

Giuseppe Puddu, Raffaele Pelorosso, Federica Gobattoni & Maria Nicolina Ripa<br />

Multi-scale analysis of carbon stocks <strong>and</strong> deforestation monitoring Case of the Eastern tropical humid forest 398<br />

of Madagascar<br />

Harifidy Rakoto Ratsimba, L.G. Rajoelison, F.M. Rabenilalana, J. Bogaert & E. Haubruge


Characterization of a Maculinea alcon population in the Alvão Natural Park (Portugal) by a mark-recapture method 404<br />

Maria Conceição Rodrigues, Patrícia Soares, José Aranha & Paula Seixas Arnaldo<br />

L<strong>and</strong>scape changes in a watershed in the southwest of Portugal 409<br />

Maria Teresa Calvão Rodrigues & Vanessa Silva<br />

<strong>L<strong>and</strong>scapes</strong> in Transition – Monitoring in Areas of L<strong>and</strong>slides 415<br />

Jorgeane Schaefer-Santos & Christel Lingnau<br />

The process of urbanization <strong>and</strong> the environment - the case of irregular occupations in the city of Ponta Grossa, 420<br />

PR, Brazil<br />

S<strong>and</strong>ra Maria Scheffer & Édina Schimanski<br />

The deforestation of loess upl<strong>and</strong>s of SE Pol<strong>and</strong> <strong>and</strong> its stages as documented by valley deposits 425<br />

(case study: Bystra river valley, Lublin Upl<strong>and</strong>)<br />

Józef Superson, Jan Reder & Wojciech Zgłobicki<br />

Impacts of changes in l<strong>and</strong> use <strong>and</strong> fragmentation patterns on Atlantic coastal forests in northern Spain 431<br />

Alberto L. Teixido, Luis G. Quintanilla, Francisco Carreño & David Gutiérrez<br />

Assessing multi-temporal l<strong>and</strong> cover changes in the Mata Nacional da Peneda Geres National Park (1995 <strong>and</strong> 2009), 437<br />

Portugal - a l<strong>and</strong> change modeler approach for l<strong>and</strong>scape spatial patterns modelling <strong>and</strong> structural evaluation<br />

Helder Viana & José Aranha<br />

Mapping invasive species (Acacia dealbata Link) using ASTER/TERRA <strong>and</strong> LANDSAT 7 ETM+ imagery 443<br />

Helder Viana & José Aranha<br />

Section 6 449<br />

Tools of l<strong>and</strong>scape assessment <strong>and</strong> management<br />

Investigation on mountain l<strong>and</strong>scape parameters on Juniper species growth (case study: Firozkooh region, Tehran) 450<br />

Nazi Avani, Hamid Jalilv<strong>and</strong> & Vahid Etemad<br />

Spatial dynamics of chestnut blight disease at the plot level using the Ripley’s K function 456<br />

João C. Azevedo, Valentim Coelho, João P. Castro, Diogo Spínola & Eugénia Gouveia<br />

The third dimension in l<strong>and</strong>scape metrics analysis applied to central Alentejo – Portugal 462<br />

Teresa Batista, Paula Mendes & Luisa Carvalho<br />

New classification <strong>and</strong> utilization of forest functions in l<strong>and</strong>scape 468<br />

Vladimir Caboun<br />

Connecting l<strong>and</strong>scape conservation <strong>and</strong> management with traditional ecological knowledge: does it matter how 474<br />

people perceive l<strong>and</strong>scape <strong>and</strong> nature<br />

Ana M. Carvalho, Margarida T. Ramos & Amélia Frazão-Moreira<br />

Identification <strong>and</strong> characterization of forest edge segments for mapping edge diversity in rural l<strong>and</strong>scapes 480<br />

Marc Deconchat, Audrey Alignier, Philippe Espy & Sylvie Ladet<br />

Visibility analysis <strong>and</strong> visual diversity assessment in rural l<strong>and</strong>scapes 486<br />

José Gaspar, Beatriz Fidalgo, David Miller, Luis Pinto & Raúl Salas<br />

L<strong>and</strong>scape runoff, precipitation variation <strong>and</strong> reservoir limnology 491<br />

A.M. Geraldes<br />

Reservoirs: Mirrors of the surrounding l<strong>and</strong>scape 496<br />

A.M. Geraldes & M.J. Boavida<br />

Using a multi-criteria approach to fit the evaluation basis of the modified 2-D cellular automaton<br />

“Pimp Your L<strong>and</strong>scape” 502<br />

Lars Koschke, Christine Fürst, Carsten Lorz, Susanne Frank & Franz Makeschin<br />

Assessing “spatially explicit” l<strong>and</strong> use/cover change models 508<br />

Jean-François Mas, Azucena Pérez Vega & Keith Clarke


Economic valuation of environmental goods <strong>and</strong> services 514<br />

Alda Matos, Paula Cabo, Isabel Ribeiro & António Fern<strong>and</strong>es<br />

L<strong>and</strong>-use management <strong>and</strong> changes in Campania Region (Southern Italy): examples from ten regional State <strong>Forest</strong>s 520<br />

A. Migliozzi, F. Cona, A. di Gennaro, A. Mingo, A. Saracino & S. Mazzoleni<br />

Naturalness <strong>and</strong> diversity of biotopes: their impact on l<strong>and</strong>scape quality-a mathematical model<br />

Maria Luiza Porto, Horst H. Wendel, Jairo J. Zocche, Gilberto G. Rodrigues, Marisa Azzolini & Rogério Both<br />

525i<br />

Application of modern, non traditional restoration methods on brown coal localities of Czech Republic 526<br />

Michal Rehor & Vratislav Ondracek<br />

Developing models <strong>and</strong> processes to aid decision support for integrated l<strong>and</strong> management in the Canadian<br />

boreal forest 532<br />

Jonathan Russell, Ellie Prepas, Gordon Putz, Daniel W. Smith & James Germida<br />

Section 7 538<br />

Management <strong>and</strong> sustainability of changing l<strong>and</strong>scapes<br />

Sustainable forest management requires multi-stakeholder governance <strong>and</strong> spatial planning: Kovdozersky state 539<br />

forest management unit in northwest Russia<br />

Per Angelstam & Marine Elbakidze<br />

Relationship between small ruminants behaviour <strong>and</strong> l<strong>and</strong>scape features in Northeast of Portugal 545<br />

Marina Castro, José Ferreira Castro & António Gómez Sal<br />

Does forest certification contribute to boreal biodiversity conservation Swedish <strong>and</strong> Russian experiences 551<br />

Marine Elbakidze, Per Angelstam, Kjell Andersson, Mats Nordberg & Yurij Pautov<br />

Impact of tree species replacement on carbon stocks in forest floor <strong>and</strong> mineral soil 557<br />

Felícia Fonseca & Tomás de Figueiredo<br />

Disentangling recent changes in forest bird ranges in Mediterranean forests (NE Spain): Assessing global change 563<br />

impacts <strong>and</strong> guiding l<strong>and</strong>scape management<br />

Assu Gil-Tena, Lluís Brotons & Santiago Saura<br />

Assessment of human <strong>and</strong> physical factors influencing spatial distribution of vegetation degradation - 569<br />

Environmental Protection Area Cachoeira das Andorinhas, Brazil<br />

Marise Barreiros Horta & Edwin Keizer<br />

Role of planted forests <strong>and</strong> trees outside forests in sustainable forest management in Iran 575<br />

E. Kouhgardi & M. Akbarzadeh<br />

Values of mangroves <strong>and</strong> its interaction with marine ecosystem 581<br />

E. Kouhgardi, E. Shakerdargah & M. Akbarzadeh<br />

The importance of Environmental Education to restore <strong>and</strong> preserve natural <strong>and</strong> cultural heritage: the case of 586<br />

Pirai da Serra – South of Brazil<br />

Edina Schimanski<br />

The certification of forest management <strong>and</strong> its contribution to the rights of workers 592<br />

Lenir Aparecida Mainardes da Silva & Edina Shimanki<br />

Role of non-wood forest products for sustainable development of rural communities in countries with a transition: 597<br />

Ukraine as a case study<br />

N. Stryamets, M. Elbakidze, P. Angelstam & R. Axelsson<br />

The regional analysis of forest management risks (by the example of Russian northern areas) 603<br />

E. Volkova<br />

Section 8 609<br />

Urban forestry in changing regions<br />

Biodiversity <strong>and</strong> recreational values in urban participatory forest planning in Finl<strong>and</strong> 610<br />

Irja Löfström, Mikko Kurttila, Leena Hamberg & Laura Nieminen


Urban tree inventory <strong>and</strong> socio-economic aspects of three villages of Ponta Grossa, PR 614<br />

Ana Carolina Rodrigues de Oliveira & Sílvia Méri Carvalho<br />

Lisbon’s public gardens, host place for world’s trees 620<br />

Isabel Silva, Elsa Isidro, Ana Luísa Soares & Francisco Moreira<br />

Section 9 626<br />

Symposia<br />

Quantifying the effects of forest fragmentation: implications for l<strong>and</strong>scape planners <strong>and</strong> 627<br />

resource managers<br />

Taking into account local people’s livelihood systems for a better management of forest fragments 628<br />

Zora Lea Urech & Jean-Pierre Sorg<br />

Measures of l<strong>and</strong>scape structure as ecological indicators <strong>and</strong> tools for conservation planning <strong>and</strong> 634<br />

forest management<br />

Multiscale analysis of l<strong>and</strong> use heterogeneity <strong>and</strong> dissimilarity as a support for planning strategies 635<br />

Emilio R. Diaz-Varela, Carlos J. Alvarez-López, Manuel F. Marey-Pérez & Pedro Álvarez-Álvarez<br />

Effects of l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong> age on species richness <strong>and</strong> biomass in a tropical dry forest 641<br />

J. Luis Hernández-Stefanoni, Juan Manuel Dupuy, Fern<strong>and</strong>o Tun-Dzul & Filogonio May-Pat<br />

Spatial gradients of l<strong>and</strong>scape metrics as an indicator of human influence on l<strong>and</strong>scape 647<br />

Evelyn Uuemaa, Ramon Reimets, Tõnu Oja, Arno Kanal & Ülo M<strong>and</strong>er<br />

L<strong>and</strong>scape assessment tools for adaptive management of tropical forested l<strong>and</strong>scapes 653<br />

Socio-economic diagnosis of a small region using an economic farming system modeling tool (Olympe). An 654<br />

approach from household to l<strong>and</strong>scape scales to assist decision making processes for development projects<br />

supporting conservation agriculture in Madagascar<br />

Eric Penot<br />

Network theory to conserve <strong>and</strong> reconnect forested l<strong>and</strong>scapes 660<br />

Connectivity loss in human dominated l<strong>and</strong>scape: operational tools for the identification of suitable habitat patches 661<br />

<strong>and</strong> corridors on amphibian’s population<br />

Samuel Decout, Stéphanie Manel, Claude Miaud & S<strong>and</strong>ra Luque<br />

L<strong>and</strong>scape genetics 667<br />

GEOME. Towards an integrated web-based l<strong>and</strong>scape genomics platform 668<br />

Stéphane Joost, Michael Kalbermatten & Nicolas Ray<br />

Road ecology: improving connectivity 673<br />

The effect of highways on native vegetation <strong>and</strong> reserve distribution in the State of São Paulo, Brazil 674<br />

Simone R. Freitas, Cláudia O. M. Sousa & Jean Paul Metzger<br />

Detecting vulnerable spots for ecological connectivity caused by minor road network in Alicante, Spain 679<br />

Beatriz Terrones, Sara Michelle Catalán, Aydeé Sanjinés, Encarnación Rico-Guzmán & Andreu Bonet<br />

A l<strong>and</strong>scape approach to sustainable forest management 685<br />

Is it possible to combine adaptation to climate change <strong>and</strong> maintaining of forest biodiversity 686<br />

Marja Kolström, Terhi Vilén & Marcus Lindner<br />

Ecosystem services from forests at the watershed scale 691<br />

<strong>Forest</strong>s in l<strong>and</strong>scapes: modelling forest l<strong>and</strong> cover patterns suitability for meeting future dem<strong>and</strong>s for l<strong>and</strong>scape 692<br />

goods <strong>and</strong> services<br />

Sónia M. Carvalho-Ribeiro & Teresa Pinto-Correia


Impacts of wildfires on catchment hydrology: results from monitoring <strong>and</strong> modeling studies in northwestern Iberia 698<br />

João Pedro Nunes, José Javier Cancelo, María Ermitas Rial, Maruxa Malvar, Filipa Tavares, Diana Vieira,<br />

Frederike Schumacher, Francisco Díaz-Fierros, António Dinis Ferreira, Celeste Coelho & Jan Jacob Keizer<br />

Management <strong>and</strong> conservation of Mediterranean forest l<strong>and</strong>scapes 704<br />

Testing the fire paradox: is fire incidence in Portugal affected by fuel age 705<br />

Paulo M. Fern<strong>and</strong>es, Carlos Loureiro, Marco Magalhães & Pedro Ferreira<br />

Interfaces <strong>and</strong> interactions between forest <strong>and</strong> agriculture in rural l<strong>and</strong>scapes 711<br />

When forests are managed by farmers: Implications of farm practices on forest management 712<br />

E. Andrieu, A. Sourdril, G du Bus de Warnaffe, M. Deconchat & G. Balent<br />

A framework for characterizing convergence <strong>and</strong> discrepancy in rural forest management in tropical <strong>and</strong> 718<br />

temperate environments<br />

D. Genin, Y. Aumeerudy-Thomas, G. Balent & G. Michon<br />

Do wooded elements in agricultural l<strong>and</strong>scape contribute to biological control in crops 724<br />

A. Ouin, M. Deconchat, P. Menozzi, C. Monteil1, L. Raison, A. Roume, J.P. Sarthou, A. Vialatte & G. Balent<br />

From ab<strong>and</strong>oned farml<strong>and</strong> to self-sustaining forests: challenges <strong>and</strong> solutions 729<br />

Harmonized measurements of spatial pattern <strong>and</strong> connectivity: application to forest habitats in the EBONE 730<br />

European Project<br />

Christine Estreguil & Giovanni Caudullo<br />

Restoration of biodiversity <strong>and</strong> ecosystem services in cropl<strong>and</strong>. Further research is needed but action is 736<br />

desperately needed<br />

José M. Rey Benayas


xiii<br />

Foreword<br />

Twenty years ago, Dr. T. Crow <strong>and</strong> Prof. B. Anko initiated a proposal to establish a new working party (WP)<br />

within the Division 8 of IUFRO, known as l<strong>and</strong>scape ecology. Back then, the field was in its most rapid growth<br />

period with many unknowns. Quite a few scholars challenged us about whether this young discipline was<br />

science. This challenge was partially important because l<strong>and</strong>scape ecology has strong components <strong>and</strong><br />

commitments to managers <strong>and</strong> policymakers. Today, l<strong>and</strong>scape ecology is matured with solid principles <strong>and</strong><br />

implementations in resource management. Many members of the WP made significant contributions to advance<br />

this field for its recognition. At the first international conference in 1990, there were a small h<strong>and</strong>ful of<br />

participants. After the announcement of this conference, over 400 abstracts were submitted. By August 6, 2010,<br />

there were at least 233 registered individuals from 44 countries.<br />

Previous bi-annual conferences had been held in the United States, Japan, Italy, Canada <strong>and</strong> China. In the last<br />

decade, the WP also paid much attention to publishing the papers presented at our bi-annual conferences.<br />

Several books <strong>and</strong> special issues based on these conferences have been published. The WP is now soliciting<br />

proposals for future conferences. Please contact any of the committee members during the conference to discuss<br />

your interests <strong>and</strong> plans. One particular effort made by conference organizers <strong>and</strong> the WP committee is to<br />

support students. We believe that student participation is vital for both the science of l<strong>and</strong>scape ecology <strong>and</strong> the<br />

growth of the WP. In 2008, we offered travel fellowships to over 28 graduate students. This year, over 20<br />

individuals received similar support.<br />

We would like to offer special thanks to Dr. Thomas Crow for his leadership since 1990. The WP has grown to<br />

have a permanent Webpage (http://research.eeescience.utoledo.edu/lees/IUFRO/), a listserv to promote<br />

communication (iufro8-l@mtu.edu), an online registration service <strong>and</strong> a committee structure that is composed of<br />

regional coordinators <strong>and</strong> liaisons with the International Association of L<strong>and</strong>scape Ecology (IALE). As of July,<br />

2010, there are 475 members in the WP database. The journal of L<strong>and</strong>scape Ecology reserves a special page for<br />

us to publish important developments from the WP. Since 2008, the WP started sponsoring summer short<br />

courses <strong>and</strong> regional workshops for researchers, students <strong>and</strong> managers. To keep our communication beyond<br />

this conference, I strongly encourage you to subscribe to a membership via our Webpage <strong>and</strong> share your<br />

experiences, new developments, personnel changes, etc. via the WP’s listserv.<br />

While growing, we face new challenges. The increasing influence from global change is, no doubt, a major one.<br />

The science of l<strong>and</strong>scape ecology can no longer be independent of the changes surrounding us. An equally<br />

important issue is from the increasing dem<strong>and</strong>s of people <strong>and</strong> intensified activities. These challenges are the<br />

primary reasons for us to have the theme of this conference as “<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>: New<br />

Frontiers in Management, Conservation <strong>and</strong> Restoration”. Through face-to-face interactions during the<br />

conference, I am very confident that everyone will be stimulated for new initiatives under this theme. However,<br />

we hope the stimulations will go beyond the conference <strong>and</strong> are translated to your daily actions after returning to<br />

your home.<br />

The WP appreciates the kindness of Instituto Politécnico de Bragança to organize this conference. We also owe a<br />

lot to the members of the Organization Committee, Scientific Committee <strong>and</strong> the Scholarship Committee for<br />

their quality work over the past 48 months. Without the financial <strong>and</strong> in-kind support of many organizations, we<br />

would not be able to have this great conference.<br />

Jiquan Chen<br />

Chair, IUFRO8.01.02<br />

Toledo, Ohio, USA, September 2010


xiv<br />

Preface<br />

This volume contains the contributions of numerous participants at the IUFRO L<strong>and</strong>scape Ecology Working<br />

Group International Conference, which took place in Bragança, Portugal, from 21 to 24 of September 2010. The<br />

conference was dedicated to the theme <strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> - New Frontiers in Management,<br />

Conservation <strong>and</strong> Restoration. The 128 papers included in this book follow the structure <strong>and</strong> topics of the<br />

conference. Sections 1 to 8 include papers relative to presentations in 18 thematic oral <strong>and</strong> two poster sessions.<br />

Section 9 is devoted to a wide-range of l<strong>and</strong>scape ecology fields covered in the 12 symposia of the conference.<br />

The Proceedings of the IUFRO L<strong>and</strong>scape Ecology Working Group International Conference register the growth<br />

of scientific interest in forest l<strong>and</strong>scape patterns <strong>and</strong> processes, <strong>and</strong> the recognition of the role of l<strong>and</strong>scape<br />

ecology in the advancement of science <strong>and</strong> management, particularly within the context of emerging physical,<br />

social <strong>and</strong> political drivers of change, which influence forest systems <strong>and</strong> the services they provide. We believe<br />

that these papers, together with the presentations <strong>and</strong> debate which took place during the IUFRO L<strong>and</strong>scape<br />

Ecology Working Group International Conference – Bragança 2010, will definitively contribute to the<br />

advancement of l<strong>and</strong>scape ecology <strong>and</strong> science in general.<br />

For their additional effort <strong>and</strong> commitment, we thank all the participants in the conference for leaving this record<br />

of their work, thoughts <strong>and</strong> science.<br />

The Editors<br />

João Carlos Azevedo<br />

Manuel Feliciano<br />

José Castro<br />

Maria Alice Pinto<br />

Bragança, Portugal, September 2010


Section 1<br />

Scaling in l<strong>and</strong>scape analysis


V. Cortes et al. 2010. Environmental drivers of benthic communities: the importance of l<strong>and</strong>scape metrics<br />

2<br />

Environmental drivers of benthic communities: the importance of<br />

l<strong>and</strong>scape metrics<br />

Rui Cortes 1 , Simone Var<strong>and</strong>as 1 , Samantha Hughes 1 , Marco Magalhães 1 & Amílcar<br />

Teixeira 2*<br />

1 CITAB, Universidade de Trás-os-Montes e Alto Douro, Portugal<br />

2 CIMO, <strong>ESA</strong>, Instituto Politécnico de Bragança, Portugal<br />

Abstract<br />

The distribution of aquatic communities is dependent on processes that act at multiplescales.<br />

This study comprised 270 samples distributed over 2 years <strong>and</strong> used a nested sampling design to<br />

estimate the variance associated with three spatial scales: basin, site <strong>and</strong> microhabitat. Habitat<br />

assessment was made using River Habitat Survey. The derived Habitat Quality Indices <strong>and</strong> the<br />

benthic composition were crossed with l<strong>and</strong>scape metrics <strong>and</strong> types of soil use, obtained from<br />

GIS data, using multiple non-parametric regressions <strong>and</strong> distance-based redundancy analysis.<br />

Invertebrate variation was mainly linked with intermediate scale (site) <strong>and</strong> l<strong>and</strong>scape metrics<br />

were the main drivers determining local characteristics. The aquatic community exhibited a<br />

stronger relationship with l<strong>and</strong>scape metrics, especially patch size <strong>and</strong> shape complexity of the<br />

dominant uses, than with habitat quality, suggesting that instream habitat improvement is a<br />

short-term solution <strong>and</strong> that stream rehabilitation must address the influence of components at<br />

higher spatial scales.<br />

Keywords: l<strong>and</strong>scape metrics, soil use, macroinvertebrates, habitat, spatial scale<br />

1. Introduction<br />

The hierarchy theory indicates that small scale physical <strong>and</strong> biological features are hierarchical<br />

nested by variables on larger spatial scales, which means that in-stream conditions are<br />

constrained <strong>and</strong> controlled by successive larger-sale factors, interacting as filters along those<br />

scales (Frissell et al. 1986; Poff, 1997). Lotic biological assemblages occurring at a given site<br />

are a subset of the potential pool of colonizers that have passed through a system of filters<br />

related to the environmental variables <strong>and</strong> their modification by human action (Boyero, 2003;<br />

Bonada et al., 2005). This is the case of geology, climate <strong>and</strong> l<strong>and</strong>scape-level factors such as<br />

l<strong>and</strong> use or vegetation patterns that have been shown to influence local habitat condition <strong>and</strong><br />

therefore the composition of benthic fauna (Roth et al. 1996; Lammert & Allan, 1999; Joy &<br />

Death, 2004). More recently, studies tended to focus on analyzing the dependence of<br />

hydromorphological characteristics on catchment level features <strong>and</strong> l<strong>and</strong>-use, in particular<br />

whether reach or catchment scale vegetation constitute suitable predictors of in-stream features<br />

(Allan, 2004; Buffagni et al., 2009; S<strong>and</strong>in, 2009). There is a strong need to develop habitat<br />

assessment strategies that integrate different complementary spatial scales from microhabitat<br />

level (including hydraulics) to the assessment of river corridor condition <strong>and</strong> surrounding l<strong>and</strong><br />

use (Cortes et al., 2009). These aspects have been already incorporated into methodologies<br />

proposed in different field surveys (e.g. Raven, 1998). There is no doubt that the spatial<br />

hierarchy of fluvial ecosystems is a crucial aspect to consider, since identifying the relationships<br />

between different levels allows associations to be made between habitat features, processes <strong>and</strong><br />

communities. This knowledge is essential for improving the implementation of appropriate<br />

*Corresponding author:<br />

Email address: amilt@ipb.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Cortes et al. 2010. Environmental drivers of benthic communities: the importance of l<strong>and</strong>scape metrics<br />

3<br />

management <strong>and</strong> monitoring measures (S<strong>and</strong>in, 2009), as the effect of multiple human pressures<br />

on aquatic habitats is spread over several spatial scales (Hughes et al., 2008).<br />

The main objective of this work was to assess the influence of environmental attributes<br />

expressed at different scales on stream communities, particularly macroinvertebrates, <strong>and</strong> to<br />

determine how the habitat descriptors are shaped by higher spatial scales, namely l<strong>and</strong>scape<br />

patterns.<br />

2. Methodology<br />

A hierarchical nested design was used for sampling benthic communities. In this study it was<br />

considered 4 catchments (Rivers Olo, Corgo, Pinhão <strong>and</strong> Tua), 15 sites distributed by the river<br />

network, 3 transects in each site <strong>and</strong> 3 micro-habitats (replicates) in each site. All catchments<br />

are located in the Douro basin (Northern Portugal) <strong>and</strong> are subjected to distinct natural<br />

conditions (such as high gradient streams ranging from 1100 m to 50 m of altitude).<br />

Furthermore, there are preserved areas, like the Olo <strong>and</strong> Tua catchments, contrasting with other<br />

areas influenced by an intensive agriculture (specially vineyards), located in the downstream<br />

sectors of rivers Corgo <strong>and</strong> Pinhão (Figure 1).<br />

Figure 1: Location of the 15 study sites along the Olo, Corgo, Pinhão <strong>and</strong> Tua rivers, all from the Douro<br />

catchment, Northern Portugal. Study sites are spread along the longitudinal axis of the main rivers, but the<br />

Tua catchment, which was only sampled in the lower section.<br />

Invertebrates sampling was made in 2006 <strong>and</strong> 2007 in these micro-habitats using surber samples.<br />

Invertebrates were identified to genus level to most of the families (except Diptera <strong>and</strong><br />

Oligochaeta).The relationship between large scale variables <strong>and</strong> the benthic fauna was assessed<br />

using both species composition <strong>and</strong> species traits. The environmental variables considered<br />

covered two levels of observation:<br />

a) At l<strong>and</strong>scape <strong>and</strong> soil use scale the data was obtained from Corine L<strong>and</strong>cover <strong>and</strong> it was<br />

considered a circle of 1km radius around each site. In the first case we used a set of metrics<br />

that can be grouped in patch metrics of density <strong>and</strong> size, edge, shape <strong>and</strong> of diversity <strong>and</strong><br />

interspersion. These metrics were further applied to each type of soil use originating a total<br />

of 48 variables (Table 1).<br />

b) At the aquatic habitat <strong>and</strong> river corridor scale it was applied the River Habitat Survey<br />

methodology (RHS - Raven et al., 1997, 1998). Ten transects or “spot checks” were made at<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Cortes et al. 2010. Environmental drivers of benthic communities: the importance of l<strong>and</strong>scape metrics<br />

4<br />

50m intervals along the 500m reach <strong>and</strong> discrete descriptions obtained (e.g. cover channel<br />

substrate, flow type, aquatic vegetation types, bank vegetation structure, artificial<br />

modifications). Continuous observations or “sweep up” along the 500m reach characterized<br />

features <strong>and</strong> modifications not described at the spot-checks (e.g. natural <strong>and</strong> man-made<br />

features, or riparian vegetation). Two habitat indices, derived from the RHS were<br />

determined: 1) Habitat Quality Assessment (HQA), that is an expression of the habitat<br />

quality (e.g., physical habitat, vegetation cover, the use of marginal l<strong>and</strong>), <strong>and</strong> 2) Habitat<br />

Modification Score (HMS) that quantifies the extent of artificialization (e.g. weirs, bank<br />

protections) along the channel.<br />

Table 1: Environmental descriptors used for describing l<strong>and</strong>scape-l<strong>and</strong> use <strong>and</strong> habitats at each of the 15<br />

sites included in the Rivers Olo, Corgo, Pinhão <strong>and</strong> Tua, The l<strong>and</strong>scape metrics were applied to the<br />

different soil use types resulting in a total of 48 l<strong>and</strong>scape variables.<br />

L<strong>and</strong>scape metrics Soil use variables Habitat descriptors<br />

PATCH DENSITY AND<br />

SIZE METRICS<br />

Number of patches<br />

Total edge<br />

Patch size st<strong>and</strong>s dev.<br />

SHAPE METRICS<br />

Mean shape index<br />

Mean patch fractal dimension<br />

Weighted mean patch fractal<br />

dimension<br />

Agriculture l<strong>and</strong>, except vineyards<br />

(area in m 2 <strong>and</strong> %)<br />

Vineyards (area: m 2 <strong>and</strong> %)<br />

Coniferous woodl<strong>and</strong> (are: m 2 ; %)<br />

Broadleaf woodl<strong>and</strong> (area: m 2 ; %)<br />

Mixed woodl<strong>and</strong> area (area: m 2 ; %)<br />

Urban area (area in m 2 <strong>and</strong> %)<br />

Scrub & shrubs (area in m 2 <strong>and</strong> %)<br />

Water surface including reservoirs<br />

<strong>and</strong> wetl<strong>and</strong>s (area in m 2 <strong>and</strong> %)<br />

ARTIFICIAL FEATURES<br />

Habitat Modification Score (HMS)<br />

HABITAT QUALITY<br />

HQA flow<br />

HQA channel<br />

HQA bank features<br />

HQA bank vegetation structure<br />

HQA point bars<br />

HQA in-stream channel vegetation<br />

HQA l<strong>and</strong> use<br />

HQA trees<br />

HQA special features<br />

A nested permutational MANOVA from resemblance matrix based on the Bray-Curtis<br />

coeficient was used to test the significance of benthic composition from the different spatial<br />

levels (catchment, site <strong>and</strong> transect). Multiple non-parametric regressions from distancebased<br />

linear models (DISTLM) were established between invertebrate taxa <strong>and</strong> the<br />

environmental variables by using the following independent variables (separately):<br />

habitat quality indices, soil use variables <strong>and</strong> l<strong>and</strong>scape metrics (Table 1). Ordination<br />

techniques using distance-based redundance analysis (dbRDA) were established between<br />

benthic fauna, but expressed as metrics sensitive to contamination <strong>and</strong> soil use <strong>and</strong> l<strong>and</strong>scape<br />

metrics. The biological metrics were extracted from Var<strong>and</strong>as & Cortes (2009) since they<br />

proved to be the most sensitive to disturbance in catchments of North Portugal (Table 2).<br />

Multivariate analyses were carried out using the package PERMANOVA for PRIMER<br />

(Anderson et al., 2008).<br />

Table 2: List of invertebrate metrics selected (Oliveira & Cortes, 2009). Acronyms are indicated in bold.<br />

Invertebrate metrics<br />

Families of Predators fP % Shredders %Shr<br />

Fam. of Ephem., Plecop., Trichop. fEPT % Scrapers %Scr<br />

Families of Swimmers fSwi % Filterers %Fil<br />

Families of Clingers fCling % Gatherers %Gath<br />

% Rheophilous %Rhe % Predators %Pred<br />

Genus of shredders gShr % Limnophilous %Lim<br />

Genus of Filterers gFil % Omnivorous %Omn<br />

Families of Gatherers fGath % organisms with branchial respiration %br<br />

% Intolerants %Int % organisms with cutaneous respiration %cr<br />

Index IBMWP % organisms with aerial respiration %ar<br />

Index FBI % organisms parasites %Par<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Cortes et al. 2010. Environmental drivers of benthic communities: the importance of l<strong>and</strong>scape metrics<br />

5<br />

Results<br />

The results of the MANOVA for benthic composition are presented on Table 3, separately for<br />

each year, <strong>and</strong> showed that site produced the significant differences for both years (p


V. Cortes et al. 2010. Environmental drivers of benthic communities: the importance of l<strong>and</strong>scape metrics<br />

6<br />

The dbRDA ordinations, grouping biological <strong>and</strong> environmental data are represented on Figure<br />

2, with separate plots for biological <strong>and</strong> environmental variables. The 1st axis reflected the<br />

longitudinal variation, where the l<strong>and</strong>scape variables linked to the natural cover (forest <strong>and</strong><br />

shrubs) define the sites located upstream <strong>and</strong> the ones associated with the vineyard influence<br />

more the lower reaches. On the first sites we may notice the presence of less disturbed<br />

communities reflected by higher values of the biotic index, EPT, shredders <strong>and</strong> intolerant fauna,<br />

whereas this pattern is replaced downstream by a dominance of organisms with branchial <strong>and</strong><br />

cutaneous respiration, belonging to trophic groups with mainly filterers <strong>and</strong> gatherers.<br />

Concerning the relation between benthic fauna <strong>and</strong> soil use it was also detected a<br />

longitudinal gradient related to soil use from natural areas (e.g. hardwood forest) to<br />

agriculture (including vineyards).<br />

Figure 2: Redundance analysis (dbRDA) between invertebrate metrics <strong>and</strong> l<strong>and</strong>scape metrics. The left<br />

diagram represents the biological metrics <strong>and</strong> on the one on the right represents the l<strong>and</strong>scape patterns.<br />

Acronyms for l<strong>and</strong>scape metrics: N number of patches; SD- patch size st<strong>and</strong>ard deviation, SI- mean<br />

shape index, SH- mean patch fractal dimension, FR- area weighted mean patch fractal dimension; the last<br />

letters represent vegetation types: VIN vineyard; AG agriculture; FF broadleaf forest; FR coniferous<br />

forest; FM mixed forest; URB urban area (see Table 2 for the biological metric codes)<br />

Discussion<br />

Many studies using environmental variables determined at different spatial levels,<br />

attempt to extract the relevant scales that lend structure to aquatic communities such as<br />

benthic macroinvertebrate assemblages (Roth et al., 1996; Lammert <strong>and</strong> Allan, 1999).<br />

Lowe et al. (2006), who made a revision on patterns <strong>and</strong> processes across multiple<br />

scales of stream-habitat organization, emphasize the fractal network structure of stream<br />

systems at a l<strong>and</strong>scape scale <strong>and</strong> mention the need to underst<strong>and</strong> how the spatial<br />

configuration of habitats within a network affect fluxes of individuals <strong>and</strong> materials.<br />

The same authors conclude that broader use of multiscale approaches to explore<br />

population <strong>and</strong> community dynamics <strong>and</strong> species-ecosystem linkages in streams will<br />

produce research results that are applicable to management <strong>and</strong> conservation challenges.<br />

This study found that l<strong>and</strong>scape metrics provided a powerful tool for assessing both<br />

macroinvertebrate dynamics, instream habitats <strong>and</strong> also the river corridor. It was also<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Cortes et al. 2010. Environmental drivers of benthic communities: the importance of l<strong>and</strong>scape metrics<br />

7<br />

found that soil use descriptors were associated with the typological functioning of the<br />

river system displayed by the longitudinal succession of benthic assemblages.<br />

References<br />

Allan J.D., 2004. <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> riverscapes: The influence of l<strong>and</strong> use on stream ecosystems.<br />

Annual Review of Ecology Evolution <strong>and</strong> Systematics, 35: 257-284.<br />

Anderson M.J., Gorley R.N. <strong>and</strong> Clarke K.R., 2008. PERMANOVA+ for PRIMER: Guide to<br />

Software <strong>and</strong> Statistical Methods. Plymouth, U.K.: Primer-E Ltd.<br />

Bonada N., Zamora-Muñoz C., Rieradevall M. <strong>and</strong> Prat N., 2005. Ecological <strong>and</strong> historical<br />

filters constraining spatial caddisfly distribution in Mediterranean rivers. Freshwater Biology,<br />

50: 81-797.<br />

Boyero L., 2003. Multiscale patterns of spatial variation in stream macroinvertebrate<br />

communities. Ecological Research, 18: 365-379.<br />

Buffagni A, Casalegno C. <strong>and</strong> Erba S., 2009. Hydromorphology <strong>and</strong> l<strong>and</strong> use at different spatial<br />

scales: expectations in a changing climate scenario for medium-sized rivers of the Western<br />

Italian Alps. Fundamental <strong>and</strong> Applied Ecology, 74: 7-25.<br />

Cortes R.M.V., Hughes S.J., Var<strong>and</strong>as S.G.P., Magalhães M. <strong>and</strong> Ferreira M.T., 2009. Habitat<br />

variation at different scales <strong>and</strong> biotic linkages in lotic systems: consequences for<br />

monitorization. Aquatic Ecology 43: 1107-1120.<br />

Frissell C.A., Liss W.J., Warren C.E. <strong>and</strong> Hurley M.D., 1986. A Hierarchical Framework for<br />

Stream Habitat Classification- Viewing Streams in a Watershed Context. Environmental<br />

Management, 10: 199-214.<br />

Hughes S.J, Ferreira M.T. <strong>and</strong> Cortes R.M.V., 2008. Hierarchical spatial patterns <strong>and</strong> drivers of<br />

change in benthic macroinvertebrate communities in an intermittent Mediterranean river.<br />

Aquatic Conservation: Marine <strong>and</strong> Freshwater Ecosystems, 18: 742-760.<br />

Joy M.K. <strong>and</strong> Death R.G., 2004. Predictive modelling <strong>and</strong> spatial mapping of freshwater fish<br />

<strong>and</strong> decapod assemblages: an integrated GIS <strong>and</strong> neural network approach. Freshwater<br />

Biology, 49: 1036-1052.<br />

Lammert M. <strong>and</strong> Allan J.D., 1999. Environmental Auditing: Assessing biotic integrity of<br />

streams: Effects of scale in measuring the influence of l<strong>and</strong> use/cover <strong>and</strong> habitat structure<br />

on fish <strong>and</strong> macroinvertebrates. Environmental Management, 23: 257–270.<br />

Lowe, W.H., Likens G.E. <strong>and</strong> Power M.E., 2006. Linking Scales in Stream Ecology. BioScience,<br />

55: 591-597.<br />

Poff N.L., 1997. L<strong>and</strong>scape filters <strong>and</strong> species traits: Towards mechanistic underst<strong>and</strong>ing <strong>and</strong><br />

prediction in stream ecology. Journal of the North American Benthological Society, 16: 391-<br />

409.<br />

Raven P.J., P. Fox J.A., Everard M., Holmes N.T.H. <strong>and</strong> Dawson F.H., 1997. River Habitat<br />

Survey: a new system for classifying rivers according to their habitat quality. In: Boon, P.J.,<br />

Howell, D.L. (eds). Freshwater quality: defining the indefinable The Stationery office,<br />

Edinburg, 215-234 pp.<br />

Raven P.J., Holmes N.T.H., Dawson F.H., Fox P.J.A., Everard M., Fozzard I.R. <strong>and</strong> Rouen K.J.,<br />

1998. River habitat quality: the physical character of rivers <strong>and</strong> streams in the UK <strong>and</strong> the<br />

Isle of Man. River Habitat Survey report no. 2, Environment Agency, Bristol.<br />

Roth N.E., Allan J.D. <strong>and</strong> Erickson D.L., 1996. L<strong>and</strong>scape influences on stream biotic integrity<br />

assessed at multiple spatial scales. L<strong>and</strong>scape Ecology, 11: 141–156.<br />

S<strong>and</strong>in L., 2009. The relationship between l<strong>and</strong>-use, hydromorphology <strong>and</strong> river biota at<br />

different spatial <strong>and</strong> temporal scales: a synthesis of seven case studies. Fundamental <strong>and</strong><br />

Applied Ecology, 174: 1-5.<br />

Var<strong>and</strong>as S.G. & Cortes R.M.V., 2009. Evaluating macroinvertebrate biological metrics for<br />

ecological assessment of streams in northern Portugal. Environmental Monitoring<br />

Assessment, DOI 10.1007/s10661-009-0996-4.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. S. de Ron et al. 2010. Habitat suitability models for two species of forest raptors in Catalonia<br />

8<br />

Habitat suitability models for two species of forest raptors in Catalonia.<br />

Methodological consequences related with different scales <strong>and</strong> data<br />

sources<br />

David Sánchez de Ron 1 , Lluís Brotons 2 , Santiago Saura 3 & José M. García del Barrio 1,4<br />

1<br />

CIFOR-INIA, Madrid, Spain<br />

2<br />

Centre Tecnològic <strong>Forest</strong>al de Catalunya, Solsona, Spain<br />

3 ETSIM-UPM Madrid, Spain<br />

4<br />

Instituto Universitario Investigación Gestión <strong>Forest</strong>al Sostenible, UVA-INIA, Spain<br />

Abstract<br />

Species distribution along a territory is a function of different environmental variables that<br />

changes in space <strong>and</strong> time. In this communication we use GIS based information from different<br />

sources <strong>and</strong> at different scales (1x1 km 2 <strong>and</strong> 10x10 km 2 ) for elaborating habitat suitability<br />

models of two forest raptors species (Buteo buteo <strong>and</strong> Accipiter gentilis) along a north-south<br />

gradient in Cataluña. The area of study has an extent of 7000 km 2 . <strong>Forest</strong> raptors species<br />

presence/absence data on 679 1x1 km 2 grid cells of the Catalan Breeding Bird Atlas, <strong>and</strong><br />

Spanish <strong>Forest</strong> Map at a scale 1:50,000 are the main information sources. Logistic regression<br />

methods have been essayed for comparison across different scales. Unexpectedly, preliminary<br />

results show no relation between variables like l<strong>and</strong> use type or diversity of habitats <strong>and</strong> raptors<br />

presence at 1x 1 km 2 , <strong>and</strong> this relationship is only significant for Buteo buteo at 10 x 10 km 2<br />

scale.<br />

Keywords: Habitat suitability models, l<strong>and</strong> uses, logistic regression, changing scales.<br />

1. Introduction<br />

Animal species related to forest habitats are dependent on habitat extension but on habitat<br />

quality as well. Advances in knowing species distribution <strong>and</strong> habitat requirements are essential<br />

for studies of population genetics, evolutionary <strong>and</strong> conservation biology, biodiversity<br />

maintenance <strong>and</strong> territorial planning, among others. Recent georeferenced information about not<br />

only species distribution in Spain, but also related to climate, lithology, changes in l<strong>and</strong> uses or<br />

historical disturbances (wild fires, floods, wind storms) at different scales are valuable data<br />

sources that contribute to the elaboration of habitat suitability models that are the basis for<br />

connectivity or habitat fragmentation analysis (Pascual-Hortal & Saura, 2008). Specialist<br />

species have been affected by reduction of habitat extension <strong>and</strong> population fragmentation (see<br />

Farhig, 2001; Lindenmayer, et al, 2003; Alderman et al, 2005) with the consequence of<br />

reduction on effective population levels. Spanish forest raptors are not the exception, <strong>and</strong> a<br />

proper knowledge of habitat requirements <strong>and</strong> habitat suitability is needed. Underlying reasons<br />

for explain mismatch between potential <strong>and</strong> real distribution of a species are scale dependent,<br />

<strong>and</strong> the explanation of this dependence is one of the goals of habitat suitability models –HSM<br />

hereafter- (Ottaviani et al, 2004; Guisan <strong>and</strong> Thuiller, 2005).<br />

In a previous paper (García del Barrio et al, 2009) we essayed HSM for two forest raptors on<br />

two forest districts of central-east Spain with presence/absence data at 10x10 km 2 scale. The<br />

aim of this paper is to compare the performance of HSM at changing scales <strong>and</strong> using different<br />

data sources. <strong>Forest</strong> raptors species chosen are goshawk (Accipiter gentilis) <strong>and</strong> common<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. S. de Ron et al. 2010. Habitat suitability models for two species of forest raptors in Catalonia<br />

9<br />

buzzard (Buteo buteo) distributed in a noth-south gradient of 140 km in Catalonia with<br />

presence/absence data at 1x1 km 2 <strong>and</strong> frequency data at 10x10 km 2 aggregation level.<br />

2. Methodology<br />

Study zone is located in Catalonia (northeast Spain). Seventy 10 x 10 km 2 (5 x 14 quadrates)<br />

units were selected along a 140 km north-south gradient. Limits were 42º 32’ N - 41º 16’N <strong>and</strong><br />

1º 25’ E – 2º 3’E. Geographic gradient ranges from Pyrenees Mountain with altitude over 3000<br />

m to Mediterranean coast at sea level.<br />

Species data came from 679 1 x 1 km 2 quadrates surveyed in the Catalan Breeding Bird Atlas<br />

1999-2002 (Estrada et al. 2004), covering less than 10 % of the study zone. L<strong>and</strong> use data came<br />

from the Spanish <strong>Forest</strong> Map at a 1:50,000 scale. <strong>Forest</strong> raptors selected were goshawk<br />

(Accipiter gentilis) <strong>and</strong> common buzzard (Buteo buteo). Goshawk is a species with a wide<br />

distribution area across Europe, inhabits several types of forest from mountain coniferous to<br />

Mediterranean evergreen at low l<strong>and</strong>s. Common buzzard is a species broadly distributed along<br />

Iberian Peninsula, less dependent than goshawk of extensive forested areas.<br />

L<strong>and</strong> uses have been reclassified following Table 1. Mean elevation was the physiographic<br />

variable selected. Mean annual precipitation, mean annual temperature <strong>and</strong> summer<br />

precipitation were the climatic variables (Gonzalo, 2007).<br />

Generalized linear models (GLZs) had been essayed; binomial-log for presence/absence data at<br />

1 x 1 km 2 level, <strong>and</strong> normal-log for frequency data at 10 x 10 km 2 level (Statistica v.7). Firstly<br />

we examined each independent variable in relation with the dependent <strong>and</strong> secondly, if there<br />

was significance of any of the independent ones, they were put together <strong>and</strong> interactions were<br />

considered in the model.<br />

3. Result<br />

The analysis at 1 x 1 km 2 resolution have not had noteworthy relationships between dependent<br />

<strong>and</strong> independent variables, neither l<strong>and</strong> uses variables nor climatic or topographic ones. Table 2<br />

shows, for the two raptors species, matches between l<strong>and</strong> uses in habitat areas <strong>and</strong> total study<br />

area. Figure 1 represents the distribution of forested l<strong>and</strong> use in three areas with the presence of<br />

the raptors <strong>and</strong> total study area. Goshawk presence is reduced to 28 of 679 quadrates (4.1 % of<br />

the effective sampled area) <strong>and</strong> main l<strong>and</strong> uses in this area are very similar to that on the total<br />

area, only with the exception of fewer incidences of artificial l<strong>and</strong> uses. Common buzzard<br />

reaches 24 % of presence (163 of 679 1 x 1 km 2 quadrates are occupied) but no significant<br />

relationships between species presence <strong>and</strong> l<strong>and</strong> use or climatic variables have been found.<br />

At 10 x 10 km 2 scale, goshawk frequency has not relevant relationships with independent<br />

variables selected, <strong>and</strong> modeling is not possible. The case of common buzzard is different<br />

because an equation that explain 36 % of deviance can be formulated<br />

F Bb = -3.223+0.516xA+0.021xF+0.001xCxF-0.005xCxA-0.009xAxF (1)<br />

F Bb : frequency of presence at 10 x 10 km 2 (nº presences/nº total)<br />

A: % of articial areas<br />

F: % of forest areas<br />

C: % of cultivated areas<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. S. de Ron et al. 2010. Habitat suitability models for two species of forest raptors in Catalonia<br />

10<br />

Equation 1 shows that final model includes only l<strong>and</strong> uses variables, not topographic or climatic<br />

ones. Two single variables (A, F) <strong>and</strong> three interactions between A, F y C slightly explain actual<br />

distribution of common buzzard in the study area (Table 3). There is no evident explanation to<br />

the incidence of interaction variables, whereas the positive effect of forest areas is obvious but<br />

the same effect of artificial in not so much explainable.<br />

4. Discussion<br />

The paper explores the use of recently sampled data of nesting birds in Catalonia in relation<br />

with main l<strong>and</strong> uses <strong>and</strong> other topographic <strong>and</strong> climatic variables. Presence/absence data at 1 x 1<br />

km 2 resolution <strong>and</strong> frequency data at 10 x 10 km 2 where used for modeling Habitat Suitability<br />

for two forest raptors species. Surprisingly, no relationships where found at the more detailed<br />

scale, <strong>and</strong> only a slight relation could be modeled (Figure 2) for the more abundant raptor<br />

species (common buzzard) at a coarser resolution level.<br />

These are not good news because high spatial resolution data of species distribution should be<br />

the basis for modeling connectivity, habitat fragmentation <strong>and</strong> other indicators relatives to<br />

species conservation. If we could not model in relation with the main territorial requirements of<br />

a species, <strong>and</strong> where they are located across the territory, it would be difficult to predict species<br />

colonization movements or habitat loss driving forces.<br />

It should be possible that the two main reasons that affect the absence of correlation found at 1 x<br />

1 km 2 resolution, is that the sampled territory is less than 10 % of total territory <strong>and</strong> l<strong>and</strong> uses<br />

distribution in species presence plots <strong>and</strong> total plots is quite similar, where no determinant for<br />

other species <strong>and</strong> other sampling areas at the same resolution levels, but there are not very much<br />

reasons to be optimist with the problem of true or false absences. Manel et al (2001) showed<br />

some problems that are not been solved in data acquisition methods, <strong>and</strong> modeling HSM with<br />

this kind of data is not so accurate.<br />

Next question is relative to habitat saturation by a target species in a territory. This relationship<br />

depends not only of potential habitat or potential niche but also of the effective niche. There are<br />

some biotic interdependence that l<strong>and</strong> uses variables or climatic <strong>and</strong> topographic ones do not<br />

reflect. Human pressure in Spain over raptor species had historically reduced their distribution<br />

areas, driving some of them near extinction. Absence of preys is other meaningful reason that<br />

we could not evaluate in these models, <strong>and</strong> for the migrant species maybe questions relative to<br />

synchronization of movements are important too.<br />

Finally, we could conclude that the use of presence/absence data at 1 x 1 km 2 resolution for<br />

two raptors species in Cataluña, in relation with l<strong>and</strong> use, climatic <strong>and</strong> tophographic variables<br />

had not improved the slight predictive power of the HSM at 10 x 10 km 2 resolution. Including<br />

variables of avian richness, potential prey’s richness or some relative could be improving the<br />

response of models but, in general, data relative to these variables are not available at broad<br />

scales.<br />

References<br />

Alderman, J., McCollin, D., Hinsley, S. A., Bellamy, P. E., Picton, P. & Crockett, R. 2005.<br />

Modelling the effects of dispersal <strong>and</strong> l<strong>and</strong>scape configuration on population distribution<br />

<strong>and</strong> viability in fragmented habitat. L<strong>and</strong>scape Ecology. 20: 857-870<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. S. de Ron et al. 2010. Habitat suitability models for two species of forest raptors in Catalonia<br />

11<br />

Estrada J., Pedrocchi V., Brotons L., Herr<strong>and</strong>o S. (eds), 2004. Atles dels ocells<br />

nidificants de Catalunya 1999-2002. Institut Català d'Ornitologia (ICO)/Lynx<br />

Edicions, Barcelona, España, 638 pp.<br />

García del Barrio, J.M., Sánchez de Ron, D., de Miguel, J., Ortega, M., Elena-Rosselló, R. 2009.<br />

Modelos de disponibilidad de hábitat aplicados a dos especies de rapaces forestales en las<br />

comarcas de Urbión y el Alto Tajo. Actas del 5º congreso <strong>Forest</strong>al Español. CD<br />

publication.<br />

Gonzalo, J. (2007).Diagnosis fitoclimática de la España Peninsular. Actualización y análisis<br />

geoestadístico aplicado. Tesis Doctoral.<br />

Guisan, A. & Thuiller, W. 2005. Predicting species distribution; offering more than simple<br />

habitat models. Ecology Letters, 8: 993-1009.<br />

Fahrig, L. 2001. How much habitat is enough Biological Conservation, 100: 65-74.<br />

Lindenmayer, D.B.,H. P. Possingham, R. C. Lacy, M. A. McCarthy <strong>and</strong> Pope M. L. 2003. How<br />

accurate are population models Lessons from l<strong>and</strong>scape-scale tests in a fragmented<br />

system. Ecology Letters 6: 41-47<br />

Manel, S., Ceri Williams, H. <strong>and</strong> Ormerod, S. J. 2001. Evaluating presence-absence models in<br />

ecology: The need to account for prevalence. Journal of Applied Ecology 38: 921-931.<br />

Ottaviani,D., G. L. Lasiano y L. Boitani. 2004. Two statistical methods to validate habitat<br />

suitability models using presence-only data. Ecological Modelling. 179: 417-443.<br />

Pascual-Hortal, L. & S. Saura. 2008. Integrating l<strong>and</strong>scape connectivity in broad-scale forest<br />

planning through a new graph-based habitat availability methodology: application to<br />

capercaillie (Tetrao urogallus) in Catalonia (NE Spain). European Journal of <strong>Forest</strong><br />

Research 127: 23-31<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. S. de Ron et al. 2010. Habitat suitability models for two species of forest raptors in Catalonia<br />

12<br />

Tables <strong>and</strong> Figures<br />

Table 1: Main l<strong>and</strong> uses in the study area (14x 5 quadrates of 10 x 10 km 2 in Catalonia)<br />

L<strong>and</strong> cover class Total Area (ha) %<br />

Tree forested (F) 441000 63<br />

Cultivated areas (C) 161000 23<br />

Grassl<strong>and</strong> <strong>and</strong> pastures ( G) 35000 5<br />

Artificial (A) 35000 5<br />

Scrub l<strong>and</strong> (S) 21000 3<br />

Others (O) 7000 1<br />

Total 70000 100<br />

Table 2: Relationships between l<strong>and</strong> uses in total sampled area <strong>and</strong> areas with presence of raptor species<br />

Total area<br />

1x1<br />

Accipiter<br />

gentilis Buteo buteo Total area<br />

10X10<br />

Accipiter<br />

gentilis<br />

Buteo<br />

buteo<br />

n %area n %area n %area n %area n %area n %area<br />

Tree forested (F) 679 62.4 28 63.2 163 66.0 70 62.6 19 60.4 56 63.2<br />

Cultivated areas (C) 679 25.5 28 29.0 163 28.1 70 22.9 19 24.7 56 23.8<br />

Artificial (A) 679 4.1 28 1.9 163 1.3 70 4.8 19 8.8 56 3.9<br />

Grassl<strong>and</strong> <strong>and</strong><br />

pastures ( G) 679 3.2 28 2.6 163 1.7 70 5.3 19 2.7 56 5.0<br />

Scrub l<strong>and</strong> (S) 679 2.9 28 2.4 163 1.8 70 2.9 19 1.7 56 2.5<br />

Mosaic with trees 679 0.7 28 0.0 163 0.3 70 0.5 19 0.6 56 0.5<br />

Other forested 679 0.6 28 0.1 163 0.2 70 0.7 19 0.7 56 0.7<br />

Body water 679 0.5 28 0.7 163 0.6 70 0.3 19 0.4 56 0.3<br />

Mosaic without<br />

trees 679 0.2 28 0.0 163 0.0 70 0.1 19 0.1 56 0.1<br />

Table 3: Generalized Linear Model statistics for common buzzard frequency at 10 x 10 km 2 scale.<br />

Df<br />

Residual Scaled P<br />

Desviance Chi 2 <strong>Change</strong> desviance P desviance F<br />

Intercept 69 3.15290 70<br />

A+F-C*A+C*F-A*F 64 2.00510 70 1.15 0.36 7.33<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. S. de Ron et al. 2010. Habitat suitability models for two species of forest raptors in Catalonia<br />

13<br />

Tree forested l<strong>and</strong> use<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Min Pc 5 Pc 25 Pc 50 Pc 75 Pc 95 Max<br />

Total area<br />

Ag<br />

Bb<br />

Figure 1: Distribution of l<strong>and</strong> use tree forested l<strong>and</strong> use on total sampled area in quadrates of<br />

1 x 1 km 2 , with presence of A. gentilis (Ag) or B. buteo (Bb). Min= Minimum, Pc= percentile, Max=<br />

Maximum.<br />

Figure 2: Real (A) <strong>and</strong> modeled distribution (B) on 10 x 10 km 2 for common buzzard. Circles in both A<br />

<strong>and</strong> B represent presence/absence data at 1 x 1 km 2 level.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Section 2<br />

Patterns <strong>and</strong> processes in changing l<strong>and</strong>scapes


B. Bożętka 2010. Recent relations between forestry <strong>and</strong> agriculture in Pol<strong>and</strong><br />

15<br />

Recent relations between forestry <strong>and</strong> agriculture in Pol<strong>and</strong> -<br />

the rural l<strong>and</strong>scape on the axis of openness <strong>and</strong> enclosure<br />

Barbara Bożętka *<br />

Gdańsk University, 4 J. Bażyńskiego Str., 80-952 Gdańsk, Pol<strong>and</strong><br />

Abstract<br />

The paper refers to a complex problem of l<strong>and</strong>scape change in Pol<strong>and</strong>. It concentrates<br />

on quality as well as quantity transformation within forestry <strong>and</strong> agriculture <strong>and</strong> their<br />

consequences for the rural l<strong>and</strong>scape. A systematic increase in a forest cover connected with a<br />

decrease in agricultural l<strong>and</strong> is observed in Pol<strong>and</strong>. The tendencies are accompanied by<br />

l<strong>and</strong>scape disturbance processes, such as settlement sprawl <strong>and</strong> intensification of agriculture.<br />

The work presented here stresses the influence of above mentioned factors on l<strong>and</strong>scape<br />

structure; it employs a broad-scale analysis <strong>and</strong> focuses on the last two decades. The issue of<br />

l<strong>and</strong>scape heterogeneity is widely investigated <strong>and</strong> regional differences have been underlined.<br />

Since the l<strong>and</strong>scape configuration experiences crucial changes, an axis of openness- enclosure<br />

<strong>and</strong> its spatial redistribution received special attention.<br />

Keywords: l<strong>and</strong>scape, structure, rural areas, reforestation, Pol<strong>and</strong><br />

1. Introduction<br />

A spatial arrangement of elements demonstrating open or enclosed character strongly<br />

determines l<strong>and</strong>scape configuration. Patterns of open <strong>and</strong> enclosed spaces organize <strong>and</strong><br />

differentiate the l<strong>and</strong>. Besides their compositional values (see Bell 2004), coexistence of openenclosed<br />

units fulfils significant ecological functions, e.g. enhancing or inhibiting material <strong>and</strong><br />

energy flows in ecosystems. The context of openness <strong>and</strong> enclosure may also underline<br />

variability of l<strong>and</strong>scapes <strong>and</strong> their dependence on a man-made impact.<br />

The Polish rural l<strong>and</strong>scape has been facing considerable changes for the last two<br />

decades, which result mainly from operation of new driving forces appearing after a political<br />

reorientation in 1989. Dem<strong>and</strong>s of a market economy altogether with a global environmental<br />

agenda, restrictions <strong>and</strong> regulations introduced by EU accession in 2004 set in motion new<br />

processes. They modify traditional l<strong>and</strong>-use patterns, influence structure, functions <strong>and</strong> values<br />

of the rural l<strong>and</strong>scape. Interestingly, the processes are well pronounced in interrelations between<br />

forestry <strong>and</strong> agriculture- two main functions held by the rural areas in Pol<strong>and</strong>.<br />

2. The Polish rural l<strong>and</strong>scape <strong>and</strong> its characteristic features<br />

Regarding a l<strong>and</strong>scape typology at a continental level, at least two types of the rural<br />

l<strong>and</strong>scape should be distinguished in Pol<strong>and</strong>: open fields <strong>and</strong> strip fields (e.g. Kostrowicki 1994,<br />

Meeus 1995). Agricultural l<strong>and</strong> has had the greatest share in the total area of the country for<br />

centuries; nowadays it equals 50.9 % (Stat.Yearb.Rep. 2010) <strong>and</strong> consists mainly of arable<br />

fields <strong>and</strong> meadows. Noteworthy, non-natural factors have dominated the evolution of the<br />

agricultural l<strong>and</strong>scape <strong>and</strong> very close relations between: 1. meadows, pastures <strong>and</strong> the river<br />

network <strong>and</strong> 2. arable l<strong>and</strong> <strong>and</strong> settlement constitute a typical structural feature of the rural<br />

* Corresponding author. Tel. 0048/58 523 65 61<br />

Email address: geobb@univ.gda.pl<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Bożętka 2010. Recent relations between forestry <strong>and</strong> agriculture in Pol<strong>and</strong><br />

16<br />

l<strong>and</strong>scape (Kostrowicki 1959). Villages frequently formed by a dispersed settlement pattern;<br />

surrounded by rather small, long <strong>and</strong> narrow fields; with the presence of diverse eco-margins<br />

<strong>and</strong> forests in a distance can give a short characteristic. Farms in Pol<strong>and</strong> are small; majority<br />

possesses their l<strong>and</strong> in separated plots, 18% of them in more than 6 parts (The Country Program<br />

of Rural…, 2009).<br />

To a lesser degree, fragmented territorial structure is also found in the Polish forest<br />

cover. Altogether with a predominance of poor biotopes, a relatively young age of trees <strong>and</strong><br />

uneven afforestation, it poses serious difficulties for maintenance of ecological functions.<br />

<strong>Forest</strong>ry is the second l<strong>and</strong>- use function in the country, the Pol<strong>and</strong>’s forest cover reached<br />

29.0% of the total area in 2009 (Stat.Yearb.Rep. 2010). This insufficient quantity is<br />

accompanied by natural <strong>and</strong> anthropogenic hazards (Liro 1998). It has been estimated that about<br />

70% of a forest cover is formed by tree st<strong>and</strong>s designated for complete felling sites. In addition,<br />

forests demonstrate a depleted species composition <strong>and</strong> a simplistic structure, where biotic<br />

components are not adjusted to existing habitats (Smykała 1993, after Rykowski 2003). In<br />

general, woodl<strong>and</strong>s have been highly fragmented <strong>and</strong> are retained now in the areas featured by<br />

the worst agricultural qualities. The greatest amounts of forests are found in the west, north,<br />

<strong>and</strong> south-east of the country (The Lubuskie Region [c. 50 %], Pomerania <strong>and</strong> Podkarpackie).<br />

3. Large-scale spatial processes <strong>and</strong> l<strong>and</strong>-use tendencies in the rural areas<br />

3.1. L<strong>and</strong>- use tendencies<br />

The Polish rural l<strong>and</strong>scape has encountered high dynamics of l<strong>and</strong>-use changes for the<br />

last twenty years. An analysis of statistical data <strong>and</strong> relevant literature (e.g. Ciołkosz, Poławski<br />

2006) allows identifying main directions of the transformation: 1. sharp decrease in meadows<br />

<strong>and</strong> pastures, 2. steady decrease in arable fields, 3. steady increase in forested areas, <strong>and</strong> 4.<br />

dynamic increase in built areas. Many authors (ibidem) stress instability of l<strong>and</strong>- use in Pol<strong>and</strong>;<br />

however, agriculture <strong>and</strong> forestry still are dominant.<br />

3.2. Agricultural areas<br />

A steady decline of agricultural l<strong>and</strong> has been noted at least since 1930s (Ciołkosz,<br />

Poławski 2006), but after the political turnover in 1989, the process accelerated again. The loss<br />

coincides with transformation of agriculture in Pol<strong>and</strong>.<br />

Incoming market requirements caused a systematic decrease in the amount of farms <strong>and</strong><br />

led to ab<strong>and</strong>onment of l<strong>and</strong> having poor soil conditions, which are being partially afforested.<br />

However, the percentage of fallow <strong>and</strong> uncultivated l<strong>and</strong> has been declining since a period<br />

2000- 2002 (11.9 % in 2000) <strong>and</strong> in 2008 accounted for 3.8 % (Environment, 2010).<br />

Simultaneously, strong dem<strong>and</strong>s for housing sites lead to designation of thous<strong>and</strong>s ha for<br />

residential areas. Shrinking of agricultural l<strong>and</strong> is accompanied by decrease in diversity of crop<br />

<strong>and</strong> breeding varieties <strong>and</strong> by strong limitation of pasture. These sharp changes are softened by<br />

environmental instruments acquired after UE accession, but polarization between highproductive<br />

areas <strong>and</strong> ab<strong>and</strong>onment of l<strong>and</strong> takes place. Contrary to arable fields, meadows <strong>and</strong><br />

pastures, the area of orchards increases <strong>and</strong> is occupied by modern, larger farms (Kulikowski<br />

2007). The Country Program of Rural Development 2007-2013 confirms further steady decline<br />

in the total agricultural l<strong>and</strong>. Its statements allow also to anticipate growing importance of l<strong>and</strong><br />

consolidation <strong>and</strong> size polarization of farms (e.g. a further loss of medium-size enterprises).<br />

3.3. <strong>Forest</strong> areas<br />

In contrast to agricultural l<strong>and</strong>, the amount of forest areas systematically grows.<br />

Dynamic post-war reforestation was strongly connected with an ownership structure- new<br />

forests emerged mostly on the state’s grounds <strong>and</strong> therefore the northern, north-western <strong>and</strong><br />

south-eastern parts of Pol<strong>and</strong> show the highest rate of a new forest cover. Distinctively, the<br />

availability of uncultivated l<strong>and</strong> during the last two decades does not correspond with the need<br />

to improve environmental <strong>and</strong> ecological functions of the rural areas. A forest cover should be<br />

especially strengthened in central <strong>and</strong> eastern parts of the country, where it suffered advanced<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Bożętka 2010. Recent relations between forestry <strong>and</strong> agriculture in Pol<strong>and</strong><br />

17<br />

fragmentation. However, supply of convenient l<strong>and</strong> is limited there mainly owing to<br />

competition of agro-environmental schemes <strong>and</strong> a growing dem<strong>and</strong> for arable fields. Therefore<br />

the planned total amount of afforestation estimated at 1.5 million ha will have to be reduced.<br />

According to the government’s plans, a forest cover in 2020 should equal 30% <strong>and</strong> after 2050<br />

33% (The Country Program for Afforestation 2003).<br />

Importantly, numerous activities aimed at improvements in forests’ quality have been<br />

undertaken. They are aimed at replacing ‘a normal forest’ scheme <strong>and</strong> mono- functional forestry<br />

by a multifunctional model (Rykowski 2003). In consequence, importance of ecological<br />

functions <strong>and</strong> renaturalisation is stressed, the share of broadleaved trees systematically grows<br />

<strong>and</strong> the age structure gets better (Raport o stanie lasów 2009). On the other h<strong>and</strong>, a quantity <strong>and</strong><br />

quality rise is accompanied by several processes initiated in farming areas, but having evident<br />

impact on forests <strong>and</strong> their ecological functions, such as settlement sprawl <strong>and</strong> intensification of<br />

agriculture. Additionally, the dem<strong>and</strong> for timber rises; requirements of energy market <strong>and</strong><br />

pressure on increase in renewable energy affect both supplies of wood <strong>and</strong> a crop structure.<br />

3. 4. L<strong>and</strong> transformation processes <strong>and</strong> l<strong>and</strong>scape degradation<br />

Although considerable environmental <strong>and</strong> ecological benefits result from a growth of a<br />

forest cover <strong>and</strong> from improvements within an ecological network, disturbing changes of<br />

l<strong>and</strong>scape structure are observed. Out of several major l<strong>and</strong> transformation processes described<br />

by Forman (2006), suburbanization, agricultural intensification, settlement sprawl <strong>and</strong><br />

reforestation are of the greatest importance in the case of the Polish countryside. First of all,<br />

contrasts between intensive <strong>and</strong> extensive l<strong>and</strong>- use become more emphasized.<br />

Expansion of vast areas featured by agricultural intensification is followed by<br />

disappearance of small biotopes <strong>and</strong> elements subdividing the l<strong>and</strong>scape <strong>and</strong> it brings with it<br />

immense l<strong>and</strong>scape homogeneity. This process is widely experienced in the northern <strong>and</strong><br />

western parts of the country. Noteworthy, processes involved in l<strong>and</strong>scape degradation often act<br />

in a synergy <strong>and</strong> spread widely. For instance, the spectacular loss of road alleys demonstrates<br />

advanced disappearance of valuable l<strong>and</strong>scape elements throughout the country (Worobiec,<br />

Liżewska 2009). Furthermore, the negative impact on the l<strong>and</strong>scape is reinforced by<br />

uncontrolled development, chaos in spatial planning <strong>and</strong> a lack of regard for l<strong>and</strong>scape values.<br />

Concerning a dimension of l<strong>and</strong>scape composition <strong>and</strong> configuration, two axes have<br />

shown high sensitivity to new conditions: order- chaos <strong>and</strong> openness- enclosure.<br />

4. Openness <strong>and</strong> enclosure of the rural l<strong>and</strong>scape<br />

Interrelations between open <strong>and</strong> enclosed space (as analysed in material physical<br />

categories) influence not only visual qualities of a l<strong>and</strong>scape, acting as one of its aesthetic<br />

descriptors (e.g. Tveit et al. 2006), but also an overall context of l<strong>and</strong>scape structure. The axis<br />

openness- enclosure is strongly connected with scale. This work, owing to a range of research is<br />

focused on a broad scale <strong>and</strong> concentrates on national or regional levels of analysis. On the base<br />

of relevant literature <strong>and</strong> the author’s own research, <strong>and</strong> concerning l<strong>and</strong>scape variability in the<br />

openness-enclosure aspect, some directions of l<strong>and</strong>scape change can be outlined (table 1).<br />

Noteworthy, the two main types of agricultural l<strong>and</strong>scape differ in their extent. Open<br />

fields are featured by a large scale, whereas strip fields units are usually formed by smaller<br />

aggregations of many similar l<strong>and</strong>scape cells. As mentioned before, owing to dem<strong>and</strong>s of<br />

intensive cereal <strong>and</strong> animal production, expansion of large fields takes place. This process is<br />

connected with field consolidation, turning meadows to arable l<strong>and</strong> <strong>and</strong> with a loss of ecomargins<br />

<strong>and</strong> semi-natural l<strong>and</strong>scape boundaries.<br />

L<strong>and</strong>scape enclosure is usually encountered in forest <strong>and</strong> semi-wild l<strong>and</strong>scapes. Both of<br />

them are connected with l<strong>and</strong> ab<strong>and</strong>onment <strong>and</strong> both can be accompanied by planned or<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Bożętka 2010. Recent relations between forestry <strong>and</strong> agriculture in Pol<strong>and</strong><br />

18<br />

unplanned reforestation. Contrary to forest <strong>and</strong> semi-wild l<strong>and</strong>scapes, an agricultural l<strong>and</strong>scape<br />

formed by large modern orchards is linked with intensive cultivation.<br />

Semi-wild l<strong>and</strong>scapes develop in two main ways. First one appears when large patches<br />

are adjoined to neighbouring forests <strong>and</strong> this usually involves occurrence of unclear boundaries.<br />

Such situation is frequently met in N <strong>and</strong> NW Pol<strong>and</strong>. Second is connected with small patches<br />

of ab<strong>and</strong>oned l<strong>and</strong>. A dispersed sequence employs tiny individual parcels or their aggregations;<br />

the sets are often separated from each other <strong>and</strong> have partly legible boundaries. This is mostly<br />

the case of an agricultural l<strong>and</strong>scape of central, eastern <strong>and</strong> southern Pol<strong>and</strong>. Parts of the<br />

agricultural l<strong>and</strong>scape transformed into wide semi-wild or forest l<strong>and</strong>scapes, to quote the<br />

example of the Bieszczady mountains, where after l<strong>and</strong> ab<strong>and</strong>onment dense associations of<br />

broadleaved trees <strong>and</strong> bushes appeared (Wanic 2009).<br />

Table 1. Tendencies of the rural l<strong>and</strong>scape transformation in Pol<strong>and</strong> in the context of openness-enclosureregional<br />

differences <strong>and</strong> main factors of change (source: author’s own research).<br />

Tendency ⇒ open l<strong>and</strong>scape<br />

Tendency ⇒ enclosed l<strong>and</strong>scape<br />

1. Open fields<br />

• increase in intensively cultivated<br />

arable fields; particularly western,<br />

north-western <strong>and</strong> central Pol<strong>and</strong><br />

• decrease in meadows <strong>and</strong> pastures<br />

(transformed into arable fields);<br />

Silesia, Wielkopolska<br />

• decrease in orchards (partly<br />

transformed into arable fields); Silesia,<br />

north- west.<br />

1. <strong>Forest</strong> l<strong>and</strong>scape<br />

• spreading out of formerly established<br />

large forests: north, north-east, northwest,<br />

western borders <strong>and</strong> south-east<br />

e.g. owing to contemporary intensive<br />

reforestation.<br />

2. Semi-wild l<strong>and</strong>scape<br />

• extensive l<strong>and</strong> ab<strong>and</strong>onment<br />

followed by natural succession <strong>and</strong><br />

later designated for forests: north,<br />

north- west, Bieszczady<br />

• l<strong>and</strong> ab<strong>and</strong>onment at a less range –<br />

individual parcels overgrowing with<br />

trees <strong>and</strong> bushes: mainly central,<br />

southern <strong>and</strong> eastern Pol<strong>and</strong><br />

3. Agricultural enclosed l<strong>and</strong>scape<br />

• increase in the amount <strong>and</strong> character<br />

of orchards (mainly Vistula Valley,<br />

central <strong>and</strong> eastern Pol<strong>and</strong>).<br />

Some of previously outlined tendencies of l<strong>and</strong>scape transformation exhibit at least one<br />

common feature: expansion of relatively huge homogeneous areas. However, organization of<br />

these areas is different. Open fields are usually formed by large patches occupied by cereals<br />

(mainly wheat, triticale, barley <strong>and</strong> lately maize), have clearly defined boundaries <strong>and</strong> a small<br />

quantity of other habitats (<strong>and</strong> inner shapes). <strong>Forest</strong> l<strong>and</strong>scapes are almost always constituted by<br />

large patches spreading out from ‘nucleus’ towards agricultural l<strong>and</strong> or constituted by dispersed<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Bożętka 2010. Recent relations between forestry <strong>and</strong> agriculture in Pol<strong>and</strong><br />

19<br />

areas trying to fill the spaces between ‘steps’ <strong>and</strong> in consequence forming a much bigger,<br />

elongated patch. Their boundaries have a mixed character, can be very sharp or gradual.<br />

Regional distribution of changes demonstrates interesting l<strong>and</strong>scape differences. A<br />

coarse- grain pattern is specific to the northern, north-western <strong>and</strong> western parts of the country.<br />

The rural area there tends to be distinctly divided into a small number of large, concentrated<br />

patches of agricultural as well as forest l<strong>and</strong>scapes- both of them featured by predominance of<br />

compositional simplicity. The relations between open <strong>and</strong> enclosed here should be stable in the<br />

near future, presumably. Central, eastern <strong>and</strong> southern parts are affected by appearance of new<br />

enclosed structures: 1. forest <strong>and</strong> ‘succession’ patches, 2. modern, very effective orchards. They<br />

systematically change l<strong>and</strong>scape configuration, to some degree introducing more diversity, but<br />

at the same time instigating potentially negative tendency, which can lead to a variable-grain or<br />

even a coarse-grain mosaic. Regarding a characteristic fragmented spatial <strong>and</strong> ownership<br />

structure, processes of l<strong>and</strong>scape change in these regions may be much more complex <strong>and</strong> much<br />

more difficult to anticipate.<br />

Hence, two important interconnected tendencies of l<strong>and</strong>scape change should be<br />

highlighted: spreading out of coarse- grained mosaics represented by large fields <strong>and</strong> forests,<br />

<strong>and</strong> shrinking of a traditional fine- grained l<strong>and</strong>scape pattern.<br />

5. Conclusions <strong>and</strong> discussion: l<strong>and</strong>scape diversity<br />

Redistribution of open <strong>and</strong> enclosed structures deeply affects l<strong>and</strong>scape heterogeneity in<br />

Pol<strong>and</strong>. Firstly, new l<strong>and</strong>scape types develop (for instance an agricultural enclosed l<strong>and</strong>scape<br />

connected with compact orchards). Secondly, considerable alterations of a l<strong>and</strong>scape mosaic<br />

appear. Open fields spread out <strong>and</strong> continuously simplify. <strong>Forest</strong> <strong>and</strong> semi-wild l<strong>and</strong>scapes<br />

spread as well, but owing to an ecological movement in forestry, they can be more diverse than<br />

before. However, influence of new patches formed by forests or other vegetation overgrowing<br />

ab<strong>and</strong>oned l<strong>and</strong> on the l<strong>and</strong>scape is difficult to assess. New elements can enrich central, eastern<br />

<strong>and</strong> southern provinces, but simultaneously, they can constitute a threat for continuity of a finegrained<br />

agricultural pattern, being able to disturb the l<strong>and</strong>scape structure <strong>and</strong> character.<br />

Operation of many processes at the same time may cause an increase in heterogeneity, but on<br />

the other h<strong>and</strong>, it may result in decline in l<strong>and</strong>scape values. Too much diversity can cause chaos<br />

(Bell 2004) <strong>and</strong> regrettably, this is a common feature of the Polish l<strong>and</strong>scape.<br />

Nevertheless, contrasts between diversity <strong>and</strong> homogeneity become stronger than they<br />

used to- vast areas of agricultural or forest monoculture generate the opposition to the traditional<br />

l<strong>and</strong>scape formed by numerous tiny units. Therefore, fragmentation of the traditional l<strong>and</strong>scape<br />

accompanied by consolidation of open fields <strong>and</strong> forest l<strong>and</strong>scapes should be emphasized.<br />

Additionally, replacement of natural by geometric shapes in l<strong>and</strong>scapes occurs (Solon 2003).<br />

Demonstrated processes vary regionally, historical differences between The Reclaimed L<strong>and</strong><br />

(the western <strong>and</strong> northern territories) <strong>and</strong> the rest of the country are reflected in l<strong>and</strong>scape<br />

evolution <strong>and</strong> in a l<strong>and</strong>scape mosaic.<br />

Some of the changes are observed globally. Palang et al. (2006), analysing l<strong>and</strong>scapes of<br />

Eastern <strong>and</strong> Central Europe noticed the passage from small-scale variability to large-scale<br />

monotony. Many authors (e.g. Richling, Solon 1996) recognized a change of rural areas that<br />

consists in two general directions of l<strong>and</strong> transformation: increase in diversity, linked for<br />

instance with ecological farming <strong>and</strong> increase in homogeneity within a l<strong>and</strong>scape in highly<br />

developed agro-ecosystems.<br />

The last two decades brought with them significant changes of l<strong>and</strong>scape structure in<br />

Pol<strong>and</strong>. The threats of advanced simplification <strong>and</strong> transition of a l<strong>and</strong>scape mosaic belong to<br />

the most considerable consequences. Sadly, the changes can lead to a great loss in the traditional<br />

l<strong>and</strong>scape.<br />

References<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Bożętka 2010. Recent relations between forestry <strong>and</strong> agriculture in Pol<strong>and</strong><br />

20<br />

Bell, S., 2004, Elements of Visual Design in the L<strong>and</strong>scape, SPON Press, London- New York;<br />

240 pp.<br />

Ciołkosz, A., <strong>and</strong> Poławski, Z., 2006, Zmiany użytkowania ziemi w Polsce w drugiej połowie<br />

XX wieku, Przegląd Geograficzny, vol. 78, No 2; 173-190.<br />

Environment 2009, Statistical Information, Central Statistical Office, Warsaw, 2010; 527 pp.<br />

Forman, R.T.T., 2006, L<strong>and</strong> Mosaics. The ecology of l<strong>and</strong>scapes <strong>and</strong> regions. Cambridge<br />

University Press, Cambridge, 632 pp.<br />

Kostrowicki, J., 1959, Badania nad użytkowaniem ziemi w Polsce, Przegląd Geograficzny, vol.<br />

31, No 3-4; 517-533.<br />

Kostrowicki, J., (ed.), 1994, Types of agriculture in Europe, 1: 2 500 000, PAN, Warsaw.<br />

Kulikowski, R., 2007, Ogrodnictwo w Polsce. Rozmieszczenie, struktura upraw i rola w<br />

produkcji rolniczej, Przegląd Geograficzny, vol. 79, No 1; 79-98.<br />

Liro, A., (Ed.) 1998, Strategia wdrażania krajowej sieci ekologicznej Econet- Polska, 1998,<br />

IUCN Pol<strong>and</strong>, Warszawa; 272 pp.<br />

Meeus, J.H.A., 1995, Pan- European l<strong>and</strong>scapes, L<strong>and</strong>scape <strong>and</strong> Urban Planning, 31 (1995);<br />

57-79.<br />

Palang, H., Printsmann, A., Konkoly- Gyuro, E., Urbanc, M., Skowronek, E., Wołoszyn, W.,<br />

2006: The forgotten rural l<strong>and</strong>scapes of Central <strong>and</strong> Eastern Europe, L<strong>and</strong>scape Ecology<br />

(2006) 21; 347-357.<br />

Raport o stanie lasów w Polsce 2008, Instytut Badawczy Leśnictwa, Warszawa, 2009; 84 pp.<br />

Richling, A., <strong>and</strong> Solon J., 1996, Ekologia krajobrazu, Wydawnictwo Naukowe PWN,<br />

Warszawa, 319 pp.<br />

Rykowski, K., 2003, Gospodarka leśna a różnorodność biologiczna [in:] Andrzejewski R., <strong>and</strong><br />

Weigle A. (Eds), Różnorodność biologiczna Polski, Narodowa Fundacja Ochrony<br />

Środowiska, Warszawa; 197-202.<br />

Solon, J., 2003, Różnorodność ponadgatunkowa- krajobrazy, [in:] Andrzejewski R., <strong>and</strong> Weigle<br />

A. (Eds), Różnorodność biologiczna Polski, NFOŚ, Warszawa; 155-160.<br />

Statistical Yearbook of The Republic of Pol<strong>and</strong>, 2009, Central Statistical Office, Warsaw, 2010.<br />

The Country Program for Afforestation, 2003, Ministry of Environment, Warsaw 2003; 107 pp.<br />

The Country Program of Rural Development 2007-2013, Ministry of Agriculture <strong>and</strong> Rural<br />

Development, Warsaw, 2009; 421 pp.<br />

Tveit, M., Ode, Å., Fry, G., 2006, Key Concepts in a Framework for Analysing Visual<br />

L<strong>and</strong>scape Character, L<strong>and</strong>scape Research, Vol. 31, No. 3; 229-255.<br />

Wanic, T., 2009: Znaczenie gatunków z rodzaju Alnus dla przekształcania gleb porolnych w<br />

Bieszczadach, [in:] Koc J. (Ed.), Mat. III Ogólnopolskiej Konf. Naukowej „Kształtowanie<br />

i Ochrona Środowiska”, Uniw. Warmińsko- Mazurski w Olsztynie, 23-25 June 2009; 228.<br />

Worobiec, K. A., Liżewska, I., (eds) 2009, Aleje przydrożne. Historia, znaczenie, zagrożenie,<br />

ochrona. Wydawnictwo „Borussia”, Kadzidłowo- Olsztyn, 271 pp.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.F. Chen et al. 2010. A physiotope-based model for ecoregions for the the nationwide ecosystem management of Japan<br />

21<br />

A physiotope-based model for ecoregions for the the nationwide<br />

ecosystem management of Japan<br />

Siew Fong Chen 1* , Tadashi Masuzawa 2 , Yukihiro Morimoto 1 & Hajime Ise 2<br />

1 Graduate School of <strong>Global</strong> Environmental Studies, Kyoto University, Japan<br />

2 Regional Environmental Planning Inc., Tokyo, Japan<br />

Abstract<br />

In this research we suggest a new ecologically-significant planning unit based on ecoregions<br />

adjusted to suit Japan's complicated geological history, <strong>and</strong> to analyze them from a l<strong>and</strong>scape<br />

ecology point of view. First, we divided Japan into four main geological regions <strong>and</strong> within<br />

them, delineated physiotope-based ecoregions using regional climate regime as main controlling<br />

factor. From a l<strong>and</strong>form-geological map overlay, we derived 7 physiotope classes <strong>and</strong><br />

characterized each geological region. We quantified the l<strong>and</strong>scape composition <strong>and</strong><br />

configuration <strong>and</strong> found that each region has different patch dynamics <strong>and</strong> mosaic complexity in<br />

which the extent of Neogene sedimentary basins <strong>and</strong> Quaternary strata deposition play a big role.<br />

Underst<strong>and</strong>ing the geotectonic-climatic-physiotope ecoregion framework of Japan is an<br />

important step in devising multi-scale, hierarchical ecosystem management.<br />

Keywords: Physiotope-based ecoregions, l<strong>and</strong>scape metrics.<br />

1. Introduction<br />

The Third National Biodiversity Strategy of Japan highlighted the biodiversity crisis <strong>and</strong> the<br />

importance of l<strong>and</strong> planning, design <strong>and</strong> ecosystem management at national spatial scale from<br />

the viewpoint of biodiversity conservation. The concept of ecoregion is suitable in devising<br />

ecologically-significant l<strong>and</strong> planning units that can be classified hierarchically <strong>and</strong> which<br />

transcends administrative boundaries. Omernik & Bailey (1997) noted that ecoregions are<br />

ecosystems of large regional extent that contain a group of geographical areas of similar<br />

functioning ecosystems, which can be delineated at different sizes <strong>and</strong> scales according to<br />

management goals. Omernik (2004) pointed out the numerous disagreements over how to<br />

delineate ecoregions e.g. disagreements on the definition of ecosystems, its complexity <strong>and</strong> its<br />

boundaries. Bailey (1996) noted the challenges of ecoregion delineation using vegetation <strong>and</strong><br />

biogeographic distribution of animal communities that constantly change due to disturbance,<br />

succession <strong>and</strong> habitat loss, therefore the importance on basing ecosystem boundaries on<br />

permanent features <strong>and</strong> dominance of one particular environmental factor. Bailey’s method sets<br />

climate as the composite, long-term, or generally prevailing weather as a primary control for<br />

ecosystem distribution at the highest level. Physiography modifies the influence on climate, <strong>and</strong><br />

has secondary effect on ecosystem differentiation.<br />

The Japanese archipelago is characterized by a narrow arc stretching on the eastern margin of<br />

the Asian continent, <strong>and</strong> is surrounded by the Japan Sea, Okhotsk Sea <strong>and</strong> Pacific Ocean which<br />

affect the different climatic regimes formed. In addition to climatic regimes, tectonic structure is<br />

equally important in ecoregion formation in Japan. Japan’s main isl<strong>and</strong>s are divided generally<br />

* Corresponding author. Tel.: +818061207975 - Fax: +81757536062<br />

Email address: siewfong.chen@at7.ecs.kyoto-u.ac.jp<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.F. Chen et al. 2010. A physiotope-based model for ecoregions for the the nationwide ecosystem management of Japan<br />

22<br />

into 2 parts: Southwest (SW) Japan <strong>and</strong> Northeast (NE) Japan by the Itoigawa-Shizuoka line,<br />

separated during the early Neogene by the subduction of the Pacific Sea Plate <strong>and</strong> the Philippine<br />

Sea Plate. A belt of Mesozoic metamorphic rocks, the Median Tectonic Line divides SW Japan<br />

into the Inner Zone <strong>and</strong> Outer Zone of SW Japan (Kimura, Hayami & Yoshida 1991). Hokkaido<br />

is separated from the main isl<strong>and</strong> of Honshu by an important zoogeographical boundary, the<br />

Blakiston Line, also known as the Tsuruga Straits.<br />

The purpose of this research is to suggest a new multi-purpose ecologically significant planning<br />

unit in Japan based on the concept of ecoregions, adjusted to suit Japan's complicated<br />

geomorphological characteristics, accretion tectonics <strong>and</strong> geological history, <strong>and</strong> finally to look<br />

at the geologic-climatic-physiographic units from a l<strong>and</strong>scape ecology point of view.<br />

2. Methodology<br />

2.1 Study site<br />

This study focuses on the main isl<strong>and</strong>s of Honshu, Hokkaido, Kyushu <strong>and</strong> Shikoku. The<br />

Ogasawara archipelago, Ryukyu archipelago <strong>and</strong> other isl<strong>and</strong>s were excluded.<br />

2.2 Delineation of ecoregions<br />

2.2.1 Delineation of ecoregions at macroscale<br />

The four major geological regions: a) Inner Zone of SW Japan, b) Outer Zone of SW Japan, c)<br />

NE Japan <strong>and</strong> d) Hokkaido were delineated using the Index Map of Fossa Magna (Kato 1992)<br />

<strong>and</strong> the Median Tectonic Line (Yoshikawa et al. 1981) <strong>and</strong> the Blakiston’s line. The Climate<br />

Regions Map of Japan by Wadachi (1974) was used to determine the boundaries of the climate<br />

regions nested within each geological zone (Fig. 1).<br />

2.2.1 Characterization of ecoregions with physiotope classes<br />

The L<strong>and</strong>form Classification map (MLIT) <strong>and</strong> Geological Classification Map (MLIT), both at<br />

1:50,000 scale <strong>and</strong> in ESRI shape format were overlaid using ArcMap 9.3 (ESRI Japan). The<br />

resulting 49 l<strong>and</strong>form-geological classes were later reclassified into 7 physiotope classes that<br />

identify accretion tectonics, tectonic relief <strong>and</strong> unique geomorphologic characteristics (Fig. 1).<br />

The resulting physiotope classes’ map was converted into ESRI GRID format <strong>and</strong> overlaid with<br />

geological regions for quantification of l<strong>and</strong>scape metrics.<br />

2.2 Quantification of l<strong>and</strong>scape metrics<br />

The raster version of FRAGSTATS (McGarigal & Marks 1994), developed by the <strong>Forest</strong><br />

Science Department, Oregon State University was used to quantify the areas, patch sizes <strong>and</strong><br />

variability, diversity, proximity <strong>and</strong> nearest neighbor indices of the physiotope classes of each<br />

geological zone. Results are in both l<strong>and</strong>scape level (Table 1) <strong>and</strong> class level (Fig. 2a-2f).<br />

3. Results<br />

3.1 Typical characteristics by geological regions<br />

3.1.1 Region A – Inner Zone of Southwest Japan<br />

Volcanoes <strong>and</strong> Volcanic L<strong>and</strong>forms <strong>and</strong> Quaternary Plains are the main physiotope classes of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.F. Chen et al. 2010. A physiotope-based model for ecoregions for the the nationwide ecosystem management of Japan<br />

23<br />

Figure 1: Proposed Physiotpe-based ecoregion map, with four main geological regions followed by<br />

regional climatic units nested within, <strong>and</strong> 7 physiotope classes for characterizing ecoregion units.<br />

this region, consisting of 25% <strong>and</strong> 22% respectively. It is a fine-grained, fragmented <strong>and</strong><br />

heterogeneous mosaic interspersed evenly across the region (Fig. 2a-2f).<br />

Analysis on the l<strong>and</strong>scape level reveals a region with the highest patch numbers, but smallest<br />

largest patch index among all regions. The highest score in interspersion <strong>and</strong> juxtaposition<br />

suggests that even though physiotope classes are small <strong>and</strong> fragmented, they remain evenly<br />

distributed across the l<strong>and</strong>scape (Table 1).<br />

3.1.2 Region B – Outer Zone of Southwest Japan<br />

Paleozoic-Mesozoic Mountain is the predominant l<strong>and</strong>scape matrix in this region with an area<br />

percentage of 57%, largest patch index of 25% (Fig. 2a & Fig. 2c). Analysis on class <strong>and</strong><br />

l<strong>and</strong>scape level show that physiotope classes has patches with the poorest connectivity <strong>and</strong> are<br />

most fragmented <strong>and</strong> isolated. Despite the scores, the patches here are strongly connected <strong>and</strong><br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.F. Chen et al. 2010. A physiotope-based model for ecoregions for the the nationwide ecosystem management of Japan<br />

24<br />

Table 1: Results of l<strong>and</strong>scape level indices by geographical regions<br />

Geological regions TA (ha) NP LPI (%) MPS (ha) PSCV (%) MPI NNCV (%) IJI (%) PR SHDI SHEI<br />

A SW Japan 12532743.75 92 7.55 136225.48 126.71 33.32 95.45 95.85 6 1.71 0.95<br />

B Outer zone Japan 3638681.25 16 25.24 227417.58 109.38 2.36 113.05 73.08 6 1.33 0.74<br />

C NE Japan 12279050.00 55 13.71 223255.45 149.08 631.52 99.75 77.24 7 1.75 0.90<br />

D Hokkaido 7794787.50 27 12.51 288695.83 88.69 18.68 120.34 83.59 6 1.59 0.89<br />

TA: Total area (ha); NP: Number of patches; LPI: Largest patch index; MPS: Mean patch size; PSCV: Patch size coefficient of<br />

variance; MPI: Mean proximity index; NNCV: Nearest-neighbor coefficient of variance; IJI: Interspersion <strong>and</strong> Juxtaposition index;<br />

PR: Patch richness; SHDI: Shannon’s Diversity Index; SHEI: Shannon’s evenness index<br />

highly homogenous, but are separated by the Hoyo Straits <strong>and</strong> the Naruto Straits (Table 1).<br />

3.1.3 Region C – Northeast Japan<br />

Tertiary Mountains <strong>and</strong> Hills <strong>and</strong> Volcanoes <strong>and</strong> Volcanic L<strong>and</strong>forms are the main physiotope<br />

classes, consisting of 30% <strong>and</strong> 22% respectively, followed by Quaternary Plains with Volcanic<br />

Ash Soil, Quaternary Plains, Mesozoic Granite Mountains <strong>and</strong> lastly Plutonic <strong>and</strong> Metamorphic<br />

Mountains (Fig. 2a). This region is characterized by the presence of an isolated Quaternary<br />

Plain with Volcanic Ash Soil patch at a 14% largest patch index (Fig. 2c). Another peculiarity is<br />

%LAND (%)<br />

60.00<br />

50.00<br />

40.00<br />

30.00<br />

20.00<br />

Thous<strong>and</strong>s<br />

MPS (ha)<br />

1800.00<br />

1600.00<br />

1400.00<br />

1200.00<br />

1000.00<br />

800.00<br />

600.00<br />

1 old<br />

2 ter<br />

3 qua<br />

4 vsh<br />

5 vol<br />

6 gra<br />

7 plu<br />

10.00<br />

0.00<br />

30.00<br />

Inner Zone SW<br />

Japan<br />

400.00<br />

200.00<br />

a) b)<br />

0.00<br />

Outer Zone SW<br />

Japan<br />

Northeast Japan<br />

Hokkaido<br />

Inner Zone SW<br />

Japan<br />

180.00<br />

Outer Zone SW<br />

Japan<br />

Northeast Japan<br />

Hokkaido<br />

25.00<br />

160.00<br />

140.00<br />

20.00<br />

120.00<br />

LPI (%)<br />

15.00<br />

PSCV (%)<br />

100.00<br />

80.00<br />

10.00<br />

60.00<br />

MPI (Index value)<br />

5.00<br />

0.00<br />

3500.00<br />

350.00<br />

3000.00<br />

300.00<br />

250.00<br />

200.00<br />

150.00<br />

100.00<br />

50.00<br />

0.00<br />

c)<br />

Inner Zone SW<br />

Japan<br />

e)<br />

Inner Zone SW<br />

Japan<br />

Outer Zone SW<br />

Japan<br />

Outer Zone SW<br />

Japan<br />

MPI = 3151.36<br />

Northeast Japan<br />

Northeast Japan<br />

Hokkaido<br />

Hokkaido<br />

NNCV (%)<br />

40.00<br />

20.00<br />

0.00<br />

180.00<br />

160.00<br />

140.00<br />

120.00<br />

100.00<br />

80.00<br />

60.00<br />

40.00<br />

20.00<br />

0.00<br />

d)<br />

Inner Zone SW<br />

Japan<br />

f)<br />

Inner Zone SW<br />

Japan<br />

Outer Zone SW<br />

Japan<br />

Outer Zone SW<br />

Japan<br />

REGIONS<br />

Northeast Japan<br />

Northeast Japan<br />

Hokkaido<br />

Hokkaido<br />

Figure 2: Comparison of l<strong>and</strong>scape indices of physiotope classes by geological regions: (a) Percentage of<br />

l<strong>and</strong>scape (%). (b) Mean patch size (ha).(c) Largest patch index (%). (d) Patch size coefficient of variance<br />

(%). (e) Mean proximity index. (f) Nearest-neighbor coefficient of variance (%). (Key to physiotope<br />

classes in Fig. 1)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.F. Chen et al. 2010. A physiotope-based model for ecoregions for the the nationwide ecosystem management of Japan<br />

25<br />

the exceedingly high mean proximity index of the Tertiary Mountains <strong>and</strong> Hills class thus<br />

making it the most connected single patch of all regions (Fig. 2e).<br />

Analysis on the l<strong>and</strong>scape level shows strong patch size heterogeneity <strong>and</strong> strongest<br />

connectivity among all regions (Table 1).<br />

3.1.4 Region D – Hokkaido<br />

Interestingly, the l<strong>and</strong> percentage of all physiotope classes in Hokkaido is almost similar to NE<br />

Japan, the only difference being the absence of Mesozoic Granite Mountains in Hokkaido.<br />

Hokkaido is characterized by fragmented coarse grain size with relatively low connectivity<br />

which is distributed unevenly across the region, particularly Paleozoic-Mesozoic Mountains <strong>and</strong><br />

Tertiary Mountains <strong>and</strong> Hills (Fig. 2a-2f).<br />

Analysis on a l<strong>and</strong>scape level shows that this region has the biggest mean patch size, lowest<br />

patch size heterogeneity, very low connectivity <strong>and</strong> most fragmented patches (Table 1).<br />

4. Discussion<br />

L<strong>and</strong>scape compostion of physiotope classes among the geological regions vary. Inner Zone of<br />

SW Japan consists of older terranes, volcanic l<strong>and</strong>forms <strong>and</strong> the highest percentage of Mesozoic<br />

granite rocks formed around 150 million years ago. Both NE Japan <strong>and</strong> Hokkaido consist<br />

mainly of newer terranes of Tertiary mountains (formed around 20 million years ago) <strong>and</strong><br />

Quaternary terranes (Fig. 2a). Jurassic accretionary prisms occupied the largest area of the pre-<br />

Neogene basement rock of SW Japan when igneous activity <strong>and</strong> regional metamorphism was<br />

most intensive. But during the Neogene era, Neogene sedimentary basins were formed only in<br />

certain parts of SW Japan while the entire of NE Japan <strong>and</strong> Hokkaido were covered with<br />

Neogene strata. During the Pleistocene-Holocene period, uppermost Pleistocene <strong>and</strong> Holocene<br />

strata formed extensive Quaternary plains in NE Japan <strong>and</strong> Hokkaido, while forming a majority<br />

of small <strong>and</strong> narrow plains in SW Japan (Kimura, Hayami & Yoshida 1991). The late<br />

Quaternary was also characterized by major changes in species distribution <strong>and</strong> composition of<br />

biotic communities (Delcourt & Delcourt 1988).The Outer Zone of SW Japan is unique both in<br />

l<strong>and</strong>scape composition <strong>and</strong> configuration, being dominated by Paleozoic-Mesozoic Mountains.<br />

In terms of physiotope class configuration, Inner Zone of SW Japan has the most complicated<br />

mosaic with small, non-contiguous, independent patches distributed all across the region. NE<br />

Japan is unique, characterized by a large, continuous patch of Tertiary Mountain <strong>and</strong> Hills class<br />

type on the Japan Sea side <strong>and</strong> a solitary Quaternary Plain with Volcanic Ash Soil in the Kanto<br />

plains (Fig. 2b-2f). Complexity of l<strong>and</strong>scape mosaics can be linked to the habitat types formed<br />

within. The dominant Tertiary mountains <strong>and</strong> hills patch of NE Japan coincide with the Fagus<br />

crenatae (broad-leaved deciduous forest) region's extent (MOE, 1989). Quaternary Plains with<br />

volcanic ash soil class type is unique because they are situated next to volcanic regions which<br />

supply the debris <strong>and</strong> being soil originating from tephra, it is among the most productive soil in<br />

the world (Shoki & Takahashi, 2002).<br />

The ecologically applicability of the results lies in ecoregion delineation of Japan. The 12<br />

climate-influenced ecoregions nested in each geological region display the significance of the<br />

Japan Sea, Okhotsk Sea <strong>and</strong> the Pacific Ocean's influences on each unit <strong>and</strong> act as the main<br />

controlling factor over the physiotope regions (Fig. 1). Even though defining the ecoregions for<br />

Japan is not an easy task <strong>and</strong> will require more extensive study, this physiotope model of<br />

ecoregions has potential as a basemap used in conjunction with environmental databases for<br />

further refinement of this study <strong>and</strong> for other research <strong>and</strong> management purposes.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.F. Chen et al. 2010. A physiotope-based model for ecoregions for the the nationwide ecosystem management of Japan<br />

26<br />

References<br />

Bailey, R.G., 1996. Ecosystem geography. New York: Springer-Verlag: 51-120<br />

Delcourt, H., & Delcourt, P., 2008. Quaternary l<strong>and</strong>scape ecology: Relevant scales in space <strong>and</strong><br />

time. L<strong>and</strong>scape Ecology, 2(1): 23-44<br />

Kato, H. 1992, Fossa Magna A masked border region separating southwest <strong>and</strong> northeast Japan.<br />

Bulletin of the Geographical Survey of Japan, 43 (1/2) :1-30<br />

Kimura, T., Hayami, I., Yoshida, S., 1991. Geology of Japan, University of Tokyo Press: 127-<br />

240<br />

McGarigal, K., Marks, B.J., 1994. FRAGSTATS: Spatial pattern analysis program for<br />

quantifying l<strong>and</strong>scape structure. Reference manual. <strong>Forest</strong> Science Department, Oregon<br />

State University, Corvallis Oregon, 134 p.<br />

Omernik, J., 2004. Perspective on the nature <strong>and</strong> the definition of ecological regions.<br />

Environmental Management, 34 suppl. 1: S27-S38<br />

Omernik, J.M. & Bailey, R.G., 1997. Distinguishing between watersheds <strong>and</strong> ecoregion.<br />

Journal of American Water Resource Association, 96178: 935-949<br />

Shoji, S., Takahashi, T., 2002. Environmental <strong>and</strong> agricultural significance of volcanic ash soils.<br />

<strong>Global</strong> Environmental Research, 6(2): 113-135<br />

Wadachi, K. ed. 1974. Kishou no jiten. Tokyo-dou Shuppan: 96<br />

Yoshikawa, T., Kaizuka, S., & Ota, Y., 1981. The L<strong>and</strong>forms of Japan. University of Tokyo<br />

Press: 50<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V.D. de Campos & S M. Carvalho 2010. Risk areas to flooding in the Hidrographical Basin of Arroio dos Pereiras<br />

27<br />

Risk areas to flooding in the Hidrographical Basin of Arroio dos<br />

Pereiras in Irati, PR, Brazil<br />

Vivian Dallagnol de Campos * & Sílvia Méri Carvalho<br />

Graduate Program in Geography - L<strong>and</strong> Management of UEPG, Brazil<br />

Abstract<br />

This study aimed to identify <strong>and</strong> map areas the risk areas to flooding in the hidrographical basin<br />

of Arroio dos Pereiras, located in Irati - PR (Brazil), identifying the most susceptible areas to<br />

accidents of hydrological origin, emphasizing the floods <strong>and</strong> co-relating the natural<br />

susceptibility derived from the natural morphodynamics of the system, where the hidrographical<br />

basin (with its inputs <strong>and</strong> outputs of energy, allied to the relief features) <strong>and</strong> the susceptibility<br />

created by human actions, mainly by the processes of urbanization, which modify the natural<br />

dynamics of the system, increasing susceptibility to the occurrence of the phenomenon <strong>and</strong><br />

creating risks to society.<br />

Word Keys: Mapping - risk areas - flooding - Irati - PR - Brazil.<br />

1. Introduction<br />

According to Tucci (2003), urbanization is one of the main constraints for the increase<br />

of the frequency <strong>and</strong> magnitude of floods, that due to changes made in the process of human<br />

occupation, such as soil sealing <strong>and</strong> deployment of infrastructure that alter the natural dynamics<br />

system. Such changes reflect in a higher runoff that leads to an increase of speed <strong>and</strong> flow of the<br />

river watercourses, potentiating the flood risk <strong>and</strong> other natural disasters linked to the dynamics<br />

of water <strong>and</strong> l<strong>and</strong> such as erosion <strong>and</strong> l<strong>and</strong>slides.<br />

Before such a reality, it is necessary to conduct studies to identify such areas at flood<br />

risk, as well as the frequency <strong>and</strong> magnitude of the event.<br />

1.1 Location <strong>and</strong> characterization of the study area.<br />

The hidrographical basin of Arroio dos Pereiras is located in Irati - PR, between the<br />

coordinates 533976/7179941 <strong>and</strong> 539906/7185815 UTM (Figure 1), comprising a total area of<br />

354 ha. It is an hidrographical basin of 3rd order, with a total of 27 river watercourses. Its main<br />

sources are located in rural areas <strong>and</strong> its riverbed runs to central urban area. Thus, the<br />

occupation seated along the hidrographical basin, eventually occupy inappropriate areas as the<br />

valleys bottoms, marginal areas <strong>and</strong> flood plains, thereby increasing the risk to flooding.<br />

The study area is included in the geological compartment of the Paraná Hidrographical<br />

Basin, with the main geological formation the Irati formation (pyrobituminous shale), with the<br />

occurrence of diabase dikes. It has an amount of unevenness to its outfall of about 110 meters,<br />

which adds up to a relief that ranges from rolling to rugged, results in areas where river flow<br />

may have a lot of energy. However, near its outfall it was modeled a floodplain quite<br />

expressive, where the urban area is settled. (Kazubek, 2006).<br />

1.2 Natural <strong>and</strong> Anthropic Constraints of the flood events.<br />

* Corresponding author. Email: vivica_pr@ibest.com.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V.D. de Campos & S M. Carvalho 2010. Risk areas to flooding in the Hidrographical Basin of Arroio dos Pereiras<br />

28<br />

According to Cunha (2005), rivers <strong>and</strong> canals overflow their riverbeds at least once<br />

every two years, in other words, when there is peak flow, triggered by concentrated rains <strong>and</strong><br />

the river begins to occupy its larger riverbed or floodplain. What makes these risk areas is<br />

human occupation, that by ignoring or disregard the natural dynamics of the drainage network<br />

provides housing in these areas <strong>and</strong> often the walls of buildings of the river banks (photo1).<br />

When the occupation of the bottom of a city valley does not consider the natural<br />

dynamic of the system, considering riverbed just the river watercourse <strong>and</strong> not respecting the<br />

larger riverbed, which is its area of flight in the events of greater flow, creates what we call risk,<br />

which is the sum of the natural susceptibility, more created susceptibility, more the vulnerability<br />

that the population is exposed (EMPLASA / SNM, 1985).<br />

Other factors that act as constraints or even aggravating of the flood events are the<br />

infrastructure projects that increase according to the urban development such as soil sealing, the<br />

reduction of green areas, changes in canals such as rectilinizations <strong>and</strong> canalizations, which,<br />

when poorly scaled, ultimately aggravate the situation, since they reduce the area occupied by<br />

the flow, blocking the water flow in the peak events, leading to leakage or even rupture of pipes<br />

<strong>and</strong> galleries.<br />

In the case studied, the hidrographical basin of Arroio dos Pereiras, the floodplain is the<br />

area of the basin that has the more consolidated urban occupation, housing commercial activities<br />

<strong>and</strong> services, in other words, it is precisely the central area of the city with a higher potential<br />

risk of material losses.<br />

2. Methodology<br />

Figure 1 - Location of the Hidrographical Basin of Arroio dos Pereiras<br />

Org: Campos, 2006.<br />

The thematic maps were generated using the software Spring 4.3 (1996), from image of<br />

QuickBird ® 2004 satellite (Irati 2006), topographic maps of the Army (SG 22-XCI-4) at<br />

1:50000 scale, base-letters of the city 1:50.000 <strong>and</strong> cadastral urban letter cadastral at 1:10,000<br />

scale.<br />

To map the risk areas they were considered the subdivision <strong>and</strong> l<strong>and</strong> use, the<br />

characteristics of the drainage network (watercourse morphology <strong>and</strong> location of flood plains),<br />

structural change <strong>and</strong> watercourse contention <strong>and</strong> historic of previous events in the local press,<br />

brigade of firemen, residents <strong>and</strong> researchers of local history. These procedures allowed to<br />

identify <strong>and</strong> map the most susceptible areas to flooding using the GIS tools.<br />

For this work we established the period from 1983 to 2006 to identify the time of<br />

recurrence of the phenomenon <strong>and</strong> they were considered only large <strong>and</strong> medium magnitude<br />

events, which have records both in the local press <strong>and</strong> in government agencies as City Hall <strong>and</strong><br />

Fire Department . Because there is not a well established civil defense, the records are few <strong>and</strong><br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V.D. de Campos & S M. Carvalho 2010. Risk areas to flooding in the Hidrographical Basin of Arroio dos Pereiras<br />

29<br />

there are not photographs that would make it possible to map more precisely the coverage areas<br />

of the flooding.<br />

Photo 1 - Margins of Arroio dos Pereiras on May 24 Avenue.<br />

Author: Campos, 2006.<br />

3. Results<br />

They were identified eight most relevant events which occurred in 1983 (2 episodes),<br />

1987, 1992, 1993, 2000, 2002 <strong>and</strong> 2004, totaling 8 events.<br />

In flood events (medium <strong>and</strong> large magnitude) all central area is reached, since it is<br />

settled on the floodplain of two hidrographical basins: Arroio dos Pereiras (our subject of study)<br />

<strong>and</strong> Antas River, being Arroio dos Pereiras tributary of Antas River.<br />

The frequency of events of greater magnitude seems to occur in a range of<br />

approximately 30 years, always in May, but for lack of documentary records from the period<br />

before 1983, it was decided to map only from 1983. It was noted during the period from 1983 to<br />

2006 that the recurrence time ranges from 40 to 10 years for the events of greater magnitude <strong>and</strong><br />

from 1 to 2 years for the events of lesser magnitude.<br />

Floods of greater magnitude (Figure 2) usually occur between the months from May to<br />

July. Of these, three major floods occurred in May (1983, 1987 <strong>and</strong> 1992).<br />

In 1983 there were two of the major floods in history, one in May <strong>and</strong> another in July,<br />

which reached the entire downtown area, where is located the center of trade <strong>and</strong> services, <strong>and</strong><br />

also different neighborhoods at different points of the city. In July of this year it was registered<br />

the highest average precipitation: 487.9 (INMET / 8th District of Meteorology). The events of<br />

1983 were compared by older residents as equal to that which occurred in early 1950 (photos 2<br />

<strong>and</strong> 3).<br />

Photo 2:03 - Flooding of the early 1950s (near 7 de Setembro Street, central area)<br />

Source: Irati (2006).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V.D. de Campos & S M. Carvalho 2010. Risk areas to flooding in the Hidrographical Basin of Arroio dos Pereiras<br />

30<br />

The flood occurred in 1993 that also had large proportions was considered atypical, not<br />

by months of occurrence (in September) which is usually rainy, but by being accompanied by a<br />

big hailstorm, which caused an average accumulation of 60 cm of ice, reaching especially the<br />

city center <strong>and</strong> within the municipality. The material losses that occurred due to this<br />

phenomenon were immense, reaching 100% in some cultures, hundreds of homes lost their<br />

roofs, fall of gas stations covering, industrial sheds, among others (Folha de Irati Newspaper,<br />

1993).<br />

Due to the magnitude of the event (heavy rain with hail), there was still clogging of the<br />

storm drainage network of rivers <strong>and</strong> the river bus by debris, trash, sediment <strong>and</strong> by the hail<br />

itself, which, coupled with the subsequent melting of the ice, led to a great flood, increasing the<br />

losses <strong>and</strong> the number of homeless. According to the local press (Folha de Irati Newspaper,<br />

1993), which covered the event, residents said in the same day (September 29, St Michael's<br />

Day), 30 years ago, the same event occurred.<br />

Figure 2 - Risk areas to flooding of great magnitude<br />

Org: Campos, 2006.<br />

Since 1993, there were few events of flooding occurred <strong>and</strong> documented: 01/2000,<br />

01/2002 <strong>and</strong> 05/2004 (local press). These were of lesser magnitude in the urban area <strong>and</strong> no<br />

record in the Fire Department. The largest losses recorded from these events were concentrated<br />

in the rural area, with losses in agricultural production, but that go beyond this study area.<br />

Structural works undertaken by the City Hall are related to reduction of the flooding<br />

frequency, although it is known the occurrence of a period of drier years, where the few<br />

episodes of more concentrated rain did not represent major risks, because the level of the river<br />

water was extremely low <strong>and</strong> the soil was very dry, with few points of flooding (Figure 3)<br />

concentrated just after the outfall of the canalization points.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V.D. de Campos & S M. Carvalho 2010. Risk areas to flooding in the Hidrographical Basin of Arroio dos Pereiras<br />

31<br />

Figure 3 - Risk Areas to Flooding of lesser Magnitude<br />

Org: Campos, 2006.<br />

4. Discussion<br />

All events covered in this survey were recorded by the City Hall, which enacted<br />

"emergency situation" for them, except the flood of 1983, where he was declared "State of<br />

Public Calamity."<br />

Floods have constituted one of the main problems for urban managers, since the<br />

solutions are costly <strong>and</strong> difficult because permeate relationships of groups with different<br />

interests <strong>and</strong> occupations for the most ancient <strong>and</strong> well established, as in the case studied.<br />

Strategize in order to alleviate the population's vulnerability to these events is the great<br />

challenge that society must be willing, starting from the underst<strong>and</strong>ing of the operation of the<br />

physical system <strong>and</strong> how the changes made do interfere in their operation, creating risks not<br />

only to environment, but mainly to society.<br />

In the areas of the hidrographical basin where the occupation is not yet consolidated,<br />

there is the possibility of seeking a rational use, as allocation of green areas for recreation,<br />

which could be used for infiltration, decreasing runoff <strong>and</strong> flood risk to downstream; non<br />

occupation of the areas around the watercourses, avoiding l<strong>and</strong>fills <strong>and</strong> embankments, among<br />

others.<br />

These are simple measures, but can effectively reduce the risk to flooding <strong>and</strong> a more<br />

harmonious relationship between man <strong>and</strong> environment.<br />

5 References<br />

Cunha, S. B. Canais Fluviais e a Questão Ambiental. p. 219-239. In: Cunha, S. B. ; Guerra, J. T.<br />

(orgs). A Questão Ambiental: diferentes abordagens. Rio de Janeiro: Bertr<strong>and</strong> Brasil,<br />

2003.<br />

EMPLASA/SNM. Disciplinamento do Uso e Ocupação do Solo com Vistas à Preservação de<br />

Inundações. In: EMPLASA/SNM. O Problema das Inundações na Gr<strong>and</strong>e São Paulo:<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V.D. de Campos & S M. Carvalho 2010. Risk areas to flooding in the Hidrographical Basin of Arroio dos Pereiras<br />

32<br />

Situação Atual e Implementação de Diretrizes Metropolitanas de Drenagem. São Paulo:<br />

EMPLASA/SNM, 1985. p. 27-42.<br />

Folha de Irati, Newspaper. Editions: 15/07/1983, 28/05 to 03/06/1992, 09 to 22/10/1993, 04 to<br />

11/06/2004.<br />

INMET, National Institute of Meteorology. 8º District – Climatological Station of Irati, 2007.<br />

Irati Hoje, Newspaper. Edition: from 14 to 20 of January, 2000.<br />

Irati. City Hall. Press Office - General Data. 2006.<br />

Kazubek, Márcio (geologist). Article: Arroio dos Pereiras – part 1. In: Jornal Hoje Centro Sul<br />

(on line edition), 2006. Available at:<br />

http://www.hojecentrosul.com.br/artigos/artigos.aspid=986. Acessado em: 20/11/2006.<br />

Spring: Integrating remote sensing<strong>and</strong> GIS by object-oriented data modelling. In: Camara G,<br />

Souza RCM, Freitas UM, Garrido J Computers & Graphics, 20: (3) 395-403, May-Jun<br />

1996.<br />

Tucci, C. E. M. Águas Urbanas. In: Tucci, C. E. M.; Bertoni, J.C. (org). Inundações Urbanas na<br />

América do Sul. Porto Alegre: Associação Brasileira de Recursos Hídricos, 2003.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N.S. Evseeva & Z.N. Kvasnikova 2010. Ecological aspects of soils deflation development in agrol<strong>and</strong>scapes<br />

33<br />

Ecological aspects of soils deflation development in agrol<strong>and</strong>scapes of<br />

the south-east of the western Siberian plain<br />

Abstract<br />

N.S. Evseeva & Z.N. Kvasnikova<br />

Tomsk State University, Russia<br />

The research of the modern processes during the cold season of the year (October-April) in the<br />

agrol<strong>and</strong>scapes of the south-western taiga area of the Western-Siberian plain has been done. The<br />

intensity of the aeol processes <strong>and</strong> the ecological <strong>and</strong> geochemical aspects of their development<br />

have been determined.<br />

Keywords: Aeol processes, taiga (thick forest) area, agrol<strong>and</strong>scape<br />

1. Introduction<br />

Most researchers regard the taiga area of the Western-Siberian plain in the western part of<br />

Siberia as the region where modern aeol processes are not practically developing. A.N. Sazhin,<br />

Ju. I. Vasilyev (2003) consider the south-east of the taiga area of the plain to be the aeol<br />

material accumulation zone. The findings obtained by the authors hold good for the natural taiga<br />

l<strong>and</strong>scapes of Western Siberia. Man’s economic activities introduce a correction for the natural<br />

course of processes <strong>and</strong> stimulate a number of them as well as aeol processes. Our research<br />

done from 1985 to 2008 shows that the modern south-eastern taiga aeol processes may be<br />

classified according to their conditions, area, <strong>and</strong> the mechanism of their development.<br />

Firstly, aeol processes are divided into natural <strong>and</strong> anthropogenic ones; secondly – they are<br />

categorized into global, regional <strong>and</strong> local ones; thirdly, they fall into destructive <strong>and</strong><br />

accumulative ones.<br />

Aeol processes do not play a great role in the relief formation of the taiga. They are represented<br />

by the near river mouth plain s<strong>and</strong> spit transfer, s<strong>and</strong> winding in uncovered places, terrace<br />

edges, flow hollows, water-shed plains as well as aeol material accumulation. The natural <strong>and</strong><br />

anthropogenic aeol processes are well-developed in arable l<strong>and</strong> areas, places of felling, oil <strong>and</strong><br />

gas extraction zones <strong>and</strong> some other areas of economic activities.<br />

Aeol dust accumulation carried by air flows over Eurasia should be called a global process. V.P.<br />

Chichagov (1999) gives examples of such a transfer. On the 5 th of May 1993 in Shizunshan<br />

(China) there occurred a transfer of fine aerosol particles from the north-west of Siberia, Central<br />

Asia <strong>and</strong> the northern part of the New L<strong>and</strong>. On the 13 th <strong>and</strong> the 14 th of April 1994, Peking got<br />

the material which had been delivered from Pol<strong>and</strong>, Sc<strong>and</strong>inavia, the mouth of the Pechorariver,<br />

Western Siberia <strong>and</strong> Central Asia. The fine material was moving across Asia by means of<br />

the north-west transfer which was constant in time. The process developing in the Western<br />

Siberia area <strong>and</strong> those ones which are connected with air mass circulation over its territory refer<br />

to the regional aeol processes.<br />

Very often they are represented by dusty storms coming to the investigated territory from<br />

Kazakhstan, Uzbekistan <strong>and</strong> the southern part of Western Siberia. The storm which occurred on<br />

27 th -28 th of April 1968, may serve as an example. At that time clouds of dust hiding the Sun<br />

hung over Tomsk <strong>and</strong> the Tomsk region as well as the Novosibirsk region. The air was greatly<br />

saturated with the light dust which was evenly distributed over the surfaces of l<strong>and</strong> <strong>and</strong><br />

buildings (Tanzybaev, Slavnina, 1975). Clouds of dust in the south-east of Western Siberia are<br />

the remains of the black storm which occurred in plowing <strong>and</strong> virgin l<strong>and</strong>s of Kazakhstan <strong>and</strong><br />

the southern part of the Western-Siberian plain (the distance from Tselinograd to Tomsk is more<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N.S. Evseeva & Z.N. Kvasnikova 2010. Ecological aspects of soils deflation development in agrol<strong>and</strong>scapes<br />

34<br />

than 1000 km) The dust composition was made up of particles less than 0,25 mm in 87,4%<br />

cases <strong>and</strong> there were particles of 0,25-0,05 mm in size in 12,6% cases. According to some<br />

approximate data, about 20 mln tons of humus, about 1 mln of nitrogen, 240 th. tons of calium<br />

<strong>and</strong> more than 60 th. tons of phosphorus were carried to the territory of the Tomsk region<br />

together with dust.<br />

The local aeol processes are those ones which develop in the range of l<strong>and</strong>scape areas on the<br />

earth’s surface (inter-river areas, terraces) which have no natural vegetation (plowl<strong>and</strong>, felling<br />

sites, the areas of oil <strong>and</strong> gas extraction).<br />

The aim of the present paper is to investigate the modern local aeol processes in the<br />

agrol<strong>and</strong>scapes which developed during the cold season of the year (October-April) from 1989<br />

to 2008. The major factors of the aeol processes development of the investigated region have<br />

been considered by N.S. Evseeva, Z.N. Kvasnikova, N.V. Osintseva, T.V. Romashova, 2001;<br />

N.S. Evseeva, V.N. Slutsky, 2005 in their works.<br />

2. Initial data <strong>and</strong> the methods of investigation<br />

In order to reveal the intensity <strong>and</strong> the dynamics of the local aeol processes development the<br />

authors have done a great deal of work. They performed repeated microscale snow survey in the<br />

vicinity of the Luchanovo station in the Tom – Yaya interriver area from 1989-2008. They also<br />

made route observations, selected snow samples but of the snow thickness <strong>and</strong> its surface as<br />

well using the support profiles; they did a three-time snow water filtering. The authors also dealt<br />

with the drying <strong>and</strong> weighing of the solid sediment, analyzed the granulometric <strong>and</strong> chemical<br />

composition of the aeol deposits. There exist various techniques of determining deflation<br />

intensivity <strong>and</strong> accumulation. They determined the deflation intensity according to the formula<br />

stated by M.E. Belgibaev (1972):<br />

h=V/S,<br />

where h is the depth of soils blowing out, m; V – the volume of the material blown out, m 3 ; S –<br />

area of dust collection, m 2 .<br />

The aeol accumulation intensity was estimated according to the aeol deposits in snow as well as<br />

aeol particles accumulation per a unit of square. According to the method developed by E.M.<br />

Lyubtsova (1994) the accumulation of aeol particles is as follows: weak (less than 50 g/m 2 ),<br />

moderate (50-100 g/m 2 ), average (100-200 g/m 2 ), strong (200-500 g/m 2 ), very strong (500-1000<br />

g/m 2 ).<br />

3. Results<br />

Aeol processes in agrol<strong>and</strong>scapes are divided into destructive <strong>and</strong> accumulative; <strong>and</strong> they are of<br />

an uneven <strong>and</strong> intermittent character. The destructive processes during the cold season of the<br />

year are represented by deflation having its centre development. Wind formed slopes of micro<strong>and</strong><br />

nano-relief elevations of a plowl<strong>and</strong> as well as furrows in the process of deep autumn<br />

plowing are subjected to deflation. Researchers observe that the Siberian soil cover is vulnerable<br />

to strong winds because the soil agregate content which is less that 1 mm reaches 40-88%, i.e. it<br />

is characterized by a strong turning to dust process. The antideflation stability of grey wood <strong>and</strong><br />

soddy-podzolic <strong>and</strong> ashen-gray plowl<strong>and</strong> soil determined by G.A. Larionov technique (1993) is<br />

not mainly high, <strong>and</strong> it varies in the range of 24-57 <strong>and</strong> 10-49. Soil particles blown out of<br />

deflation centres are carried by winds over various distances <strong>and</strong> deposited on plowl<strong>and</strong> snow<br />

surface, in shelterbelts located not far from the plowing areas.<br />

Besides, fine particles which are carried by windy flows from other regions are accumulated in<br />

snow. According to some observations made from 2000-01, 2002-03, 2003-04 <strong>and</strong> 2004-05,<br />

their weak development was seen in 2005-06, 2006-07 (28%), <strong>and</strong> their average development<br />

occurred in 2007-08 (1%).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N.S. Evseeva & Z.N. Kvasnikova 2010. Ecological aspects of soils deflation development in agrol<strong>and</strong>scapes<br />

35<br />

The average depth of soils blowing out varies from 0,1 mm up to 0,4 mm. The amount of aeol<br />

deposits accumulation during the observation period changed in the plowl<strong>and</strong> from 2,5-2,7 g/m 2<br />

(2005-06), up to 824 g/m 2 during the cold season of the year in 2002-2003.<br />

In should be noted that in the cedar forest bordering on the Luchanovo area from the east less<br />

than 18,8 g/m 2 of fine grained soil was gathered in snow thickness during the first stage. In the<br />

season of snow melting the intensity of aeol processes is high <strong>and</strong> intermittent which is due to<br />

air <strong>and</strong> soil temperature change, wind velocity, snowfall, etc. Selecting samples from the snow<br />

surface using the centre profiles shows that a great amount of aeol material can be accumulated<br />

in plowl<strong>and</strong>: from 0,83 g/m 2 to 224,5 g/m 2 , <strong>and</strong> in the cedar forest – from 0,285 g/m 2 to 1,0<br />

g/m 2 . When strong snow storms occur, up to 236 g/m 2 of fine grained soil is gathered within<br />

twenty four hours. The granulometric composition of aeol deposits is varied enough but dusty<br />

particles prevail: dust-up to 90%, fine grained s<strong>and</strong>-up to 21%, silt – 30%.<br />

Deflation <strong>and</strong> accumulation have a great effect on the ecological <strong>and</strong> geochemical processes in<br />

taiga agrol<strong>and</strong>scape territory. They cause the acceleration of natural <strong>and</strong> technogenic biophyle<br />

elements <strong>and</strong> humus migration, redistribution of the aeol material mass within a plowl<strong>and</strong> <strong>and</strong><br />

the near – by areas, accumulation of chemical elements <strong>and</strong> heavy metals as well. The<br />

macroelements essential to agricultural plants are removed <strong>and</strong> redistributed from deflation<br />

centres. Humus content in aeol deposits reaches 3,5%, N (nitrogen) up to 0,62%, P (phosphorus)<br />

– 0.56%, Cu (copper) – 95 g/t, Pb (plumbum = lead) – 16 g/t, Zn (zink) up to 68 g/t, Ba<br />

(barium) – 860 g/t, V (vanadium) – 146 g/t, etc.<br />

The study of the modern aeol processes during the cold season of the year in the agrol<strong>and</strong>scapes<br />

of the south-eastern Western Siberian taiga zone has shown:<br />

1). their development is of a cyclic character which may be well observed in the course of 2-3 or<br />

5-6 summer cycles. From 1988-89, 2000-04 aeol processes were extremely active during the<br />

cold season of the year.<br />

2). aeol processes in a plowl<strong>and</strong> affect the fertility of soils changing its mechanical, physical <strong>and</strong><br />

chemical properties.<br />

References<br />

Sazhin A.N., Vasilyev Ju.I., 2003. Geographical Regularities of Modern Deflation in the<br />

Eastern Europe <strong>and</strong> Western Siberian Steppes. Geomorphology, N1, 79-82.<br />

Chichagov V.P. 1999. Aeol Relief of Eastern Mongolia. The Institute of Geography of the<br />

Russian Academy of Sciences, 269 p.<br />

Tanzybaev M.G., Slavnina T.P. 1975. Extraordinary Phenomenon in the Nature of the Tomsk<br />

Region. Glacioclimatology of Western Siberia. Leningrad, Vol. 9, 155-160.<br />

Evseeva N.S., Kvasnikova Z.N, Osintseva N.V., Romashova T.V., 2001. Field Research of<br />

Soils Deflation of the Tom-Yaya Inter-River Area During the Cold Season of the Year.<br />

Ecological Risk: Proceedings of the 2 d All-Russian Conference. Irkutsk, 256-258.<br />

Evseeva N.S., Slutsky V.N., 2005. Climatology Factor of Aeol Processes Development in the<br />

South-East of the Western-Siberian plain. Geography <strong>and</strong> Natural Resources, N4, 75-79.<br />

Larionov G.A. 1993. Erosion <strong>and</strong> Soils deflation: the Major Regularities <strong>and</strong> Quantitative<br />

Estimation Moscow: the Moscow University Publishing Centre, 200 p.<br />

Belgibaev M.E. 1972. Natural Conditions of Soils Deflation <strong>and</strong> Soil <strong>and</strong> Erosion Division into<br />

Districts of the North-Turgai Plain. Synopsis of the Thesis of the C<strong>and</strong>idate of the<br />

Geographical Sciences. Alma-Ata, 22 p.<br />

Lyubtsova E.M. 1994. Aeol Migration of a Substance <strong>and</strong> Its Role in the Fluorine Distribution<br />

in the Southern L<strong>and</strong>scape of Minusinsk Depression. Geography <strong>and</strong> Natural Resources,<br />

N2, 86-91.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Frank et al. 2010. A regionally adaptable approach of l<strong>and</strong>scape assessment using l<strong>and</strong>scape metrics<br />

36<br />

A regionally adaptable approach of l<strong>and</strong>scape assessment using<br />

l<strong>and</strong>scape metrics within the 2D cellular automaton “Pimp your<br />

l<strong>and</strong>scape”<br />

Susanne Frank * , Christine Fürst, Carsten Lorz, Lars Koschke & Franz Makeschin<br />

Institute for Soil Science <strong>and</strong> Site Ecology, TU Dresden, Germany<br />

Abstract:<br />

We propose an easily applicable <strong>and</strong> regionally adaptable approach to quantify ecological<br />

functioning at l<strong>and</strong>scape level using l<strong>and</strong>scape metrics within the modified 2D cellular<br />

automaton Pimp your l<strong>and</strong>scape. In addition to a basic evaluation of l<strong>and</strong> use types, a<br />

consideration of l<strong>and</strong>scape patterns is essential to evaluate ecological functions <strong>and</strong> services.<br />

Starting point of the here presented approach is the aggregation of l<strong>and</strong> use types according to<br />

the degree of hemeroby. With regard to the concept of habitat connectivity, a cost- distanceprocedure<br />

allows identifying functionally connected natural areas in a next step. Further<br />

indicators to assess ecological functioning were taken into account. Intensity of l<strong>and</strong> use, habitat<br />

connectivity <strong>and</strong> habitat variety can be quantified using here presented set of metrics. We found<br />

that the assessment of l<strong>and</strong>scape structure related ecosystem services requests a combination of<br />

l<strong>and</strong>scape metrics on the one h<strong>and</strong> <strong>and</strong> scientific <strong>and</strong> normative assumptions on the other.<br />

Keywords: l<strong>and</strong>scape metrics, Pimp Your L<strong>and</strong>scape, biodiversity, habitat connectivity,<br />

l<strong>and</strong>scape fragmentation<br />

1. Introduction<br />

Numerous adaptation strategies concerning effects of Climate <strong>Change</strong> on agriculture <strong>and</strong><br />

forestry have been developed during the last decades (e.g. Rosenberg 1992; Molden 2007;<br />

Zomer et al. 2008; Allen et al. 2010). Also, various studies on effectiveness of such measures<br />

took place (e.g. Goria <strong>and</strong> Gambarelli 2004; Adger et al. 2007; Van den Berg <strong>and</strong> Feinstein<br />

2009). Visualization tools that support integrated l<strong>and</strong>scape planning are still rare. Therefore,<br />

the interactive planning tool Pimp Your L<strong>and</strong>scape was developed (Fuerst et al. 2008). It allows<br />

both, evaluation <strong>and</strong> visualization of l<strong>and</strong> use scenarios. This web-based tool supports multicriteria<br />

decision making <strong>and</strong> participatory processes in l<strong>and</strong> use management at l<strong>and</strong>scape level.<br />

It enables even the inexperienced user to design a l<strong>and</strong>scape by mouse click. Basic information<br />

of l<strong>and</strong> use classification is provided by the European wide available data set Corine L<strong>and</strong>cover<br />

2000 (CLC 2000). To evaluate complex interactions between l<strong>and</strong> use types (LUTs), it is<br />

necessary to divide the l<strong>and</strong>scape continuum into spatially distinct units, which can interact <strong>and</strong><br />

communicate. To reflect complex spatial interactions, a modified 2D cellular automaton with<br />

Moore-neighborhood was used as development basis for Pimp Your L<strong>and</strong>scape (Fuerst et al.<br />

2008). So far, l<strong>and</strong>scape structure related aspects of l<strong>and</strong>scape ecology, such as habitat<br />

connectivity <strong>and</strong> l<strong>and</strong>scape fragmentation cannot be appraised with the cellular automaton<br />

approach. In consequence, an evaluation procedure has to be developed to take l<strong>and</strong>scape<br />

structure into account. Within Pimp Your L<strong>and</strong>scape, several ecosystem services are assessed<br />

* Corresponding author. Tel.: +49 (0)35203 38-31377 - Fax: -31388<br />

Email address: Susanne.Frank@tu-dresden.de<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Frank et al. 2010. A regionally adaptable approach of l<strong>and</strong>scape assessment using l<strong>and</strong>scape metrics<br />

37<br />

such as economic wealth, ecological functioning <strong>and</strong> l<strong>and</strong>scape aesthetics. L<strong>and</strong>scape evaluation<br />

is realized at two assessment levels. The first level (a) provides an indicator based evaluation of<br />

LUT- values with regard to each ecosystem service on a relative scale from 0 to 100. The<br />

relative scale addresses the problem to provide a basis for comparing changes of various<br />

ecosystem services, which each are expressed by different units <strong>and</strong> address different orders of<br />

magnitudes. Regarding cell values, the LUT-values are corrected by integrating cell specific<br />

attributes (soil, climate, topography) <strong>and</strong> the cell specific environment (neighbored LUT).<br />

The here presented paper documents a second evaluation level (b). The assessment of l<strong>and</strong>scape<br />

patterns using l<strong>and</strong>scape metrics (LMs) will provide the possibility to correct the result achieved<br />

for the ecosystem service “ecological functioning” with regard to superior l<strong>and</strong>scape structure<br />

aspects. Aim of our study is to develop a scientifically <strong>and</strong> normative based system facilitating<br />

quantification <strong>and</strong> assessment of the ecological functioning at l<strong>and</strong>scape level.<br />

2. Conceptual approach<br />

For the usage of LMs as corrective for the ecological functioning of a l<strong>and</strong>scape, relations<br />

between l<strong>and</strong>scape mosaic, ecological processes <strong>and</strong> ecosystem services need to be identified.<br />

L<strong>and</strong>scape structure represents the interface between the l<strong>and</strong> use that is influenced by natural<br />

<strong>and</strong> cultural compounds on the one h<strong>and</strong> <strong>and</strong> ecological <strong>and</strong> functional l<strong>and</strong>scape properties on<br />

the other. Hence, quantification of l<strong>and</strong>scape structure considering composition <strong>and</strong><br />

configuration of patches is of fundamental importance for an assessment of ecosystem services<br />

(Zebisch et al. 2004).<br />

LMs can be calculated at three spatial levels: patch level, class level <strong>and</strong> l<strong>and</strong>scape level<br />

(McGarigal <strong>and</strong> Marks 1995). To ensure an adequate quantification of the considered processes,<br />

we mainly referred to class level. An aggregation of LUTs provides not only a spatial but also a<br />

functional reference. This classification approach implies three functional groups:<br />

(a) unsealed open space, e.g. Agricultural areas, <strong>Forest</strong> <strong>and</strong> semi natural areas, Wetl<strong>and</strong>s;<br />

(b) sealed areas, e.g. artificial areas, streets, rail tracks;<br />

(c) degree of hemeroby.<br />

Especially (c) is of importance because it considers function aspects of the l<strong>and</strong>scape structure.<br />

Six degrees of hemeroby are distinguished (Table 1). Each LUT was assigned to one degree of<br />

hemeroby. Additionally, a superior aggregation into “natural” <strong>and</strong> “not natural” LUTs became<br />

necessary to assess certain ecological parameters such as habitat connectivity. As no<br />

“ahemerobe” LUTs are located in Europe (Steinhardt et al. 1999), this hemeroby class was not<br />

considered for calculations.<br />

Table 1: Classification of the human impact on ecosystems <strong>and</strong> the according degree of hemeroby <strong>and</strong><br />

naturalness (Blume <strong>and</strong> Sukopp 1976 -modified)<br />

LMs describing the intensity of l<strong>and</strong> use <strong>and</strong> the habitat connectivity, which are two of three<br />

main evaluation parameters, are based on the hemeroby classification. The degree of hemeroby<br />

is used for the Hemeroby Index. Habitat connectivity is based on the sub-classification into<br />

“natural” <strong>and</strong> “not natural” LUTs. The third main indicator, biodiversity, is based on two<br />

indices measuring the spatial diversity of a l<strong>and</strong>scape. Figure 1 gives an overview of LMs,<br />

indicators <strong>and</strong> main indicators that are used to evaluate the ecological functioning.<br />

Figure 1: Hierarchical assessment scheme for the ecosystem service “ecological functioning”<br />

3. Assessment of the ecological value<br />

Figure 1 highlightens that intensity of l<strong>and</strong> use, habitat connectivity <strong>and</strong> biodiversity are used as<br />

main factors for evaluating the effect of the l<strong>and</strong>scape structure on ecological functioning.<br />

Intensity of l<strong>and</strong> use <strong>and</strong> biodiversity are integrated into the evaluation by the use of “ecological<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Frank et al. 2010. A regionally adaptable approach of l<strong>and</strong>scape assessment using l<strong>and</strong>scape metrics<br />

38<br />

linking matrices” that allow for combining various LMs including their mutual impact (Bastian<br />

<strong>and</strong> Schreiber 1999). Figure 2 illustrates the methodological approach. Values of the considered<br />

LMs are assigned to five classes. “Hemeroby” that quantifies the degree of naturalness is one<br />

factor for the determination of the intensity of l<strong>and</strong> use (left part of Figure 2). For the<br />

investigation of the second factor “l<strong>and</strong>scape fragmentation”, two LMs get interlinked. The<br />

Effective Mesh Size (M eff ) is a LM measuring the mean area of unsealed open area. A linkage<br />

with the Total Core Area of natural LUTs yields a value of “l<strong>and</strong>scape fragmentation” (right<br />

part of Figure 2). The final value of the “Intensity of L<strong>and</strong> Use” can then be read off the linking<br />

matrix (bottom part of Figure 2).<br />

Figure 2: Exemplary application of ecological linking matrices for assessing the intensity of l<strong>and</strong> use<br />

For calculating intensity of l<strong>and</strong> use <strong>and</strong> biodiversity, the same procedure is applied. For<br />

considering different aspects of biodiversity at l<strong>and</strong>scape level we considered two spatial<br />

aspects. Shannon’s Diversity Index (SHDI) that reflects the compositional component (number<br />

of patch types) as well as the structural component (distribution of classes) of l<strong>and</strong>scapes, was<br />

taken into account. Additionally, in order to quantify the spatial configuration of patch types,<br />

the Interspersion <strong>and</strong> Juxtaposition Index (IJI) was chosen. Interlinking these two indices within<br />

a linking matrix, statements on the variety of habitats can be made. For quantifying the third<br />

factor habitat connectivity, the Cost Distance Method is used. This method allows measuring<br />

functional connectivity of “natural” habitats (Zebisch et al. 2004). The basic assumption is that<br />

“ecological costs” increase with increasing degree of hemeroby of raster cells that need to be<br />

crossed for reaching other natural areas. The Moving-Window-method was then applied to<br />

identify natural areas that might serve as “stepping stones” between core areas. The value of<br />

habitat connectivity ranges from -10 to 10 in 5-point-steps (Table 2).<br />

Table 2: Evaluation of the habitat connectivity<br />

The normative background, which sets thresholds of LMs beyond expert knowledge, is given by<br />

German laws <strong>and</strong> strategies (e.g. BMU 2007; BNatSchG 2010). Thresholds were set following<br />

development targets. For example the degree of habitat connectivity m<strong>and</strong>atory shall be at least<br />

10 % (§20(1) BNatSchG 2010). As ecologists criticize this value being too low (SRU 2000;<br />

Krüsemann 2006), a positive evaluation starts at a percentage of >15 % of habitat connectivity.<br />

Summarizing the results of all three factors gives the figure to which the preliminarily<br />

calculated value for ecological functioning in Pimp Your L<strong>and</strong>scape should be increased or<br />

reduced. It ranges from - 30 to + 30. Due to the relative evaluation scale of 0 to 100, the impact<br />

of LMs as superior evaluation tool within Pimp Your L<strong>and</strong>scape is estimated to be significant.<br />

4. Conclusion <strong>and</strong> outlook<br />

Concerning l<strong>and</strong>scape panning, LMs within Pimp Your L<strong>and</strong>scape provide an assessment tool<br />

for the evaluation of functions <strong>and</strong> services that cannot easily be determined. The combination<br />

of ecological linking matrices <strong>and</strong> traits of an additive model provided the possibility to<br />

aggregate a large number of l<strong>and</strong>scape characteristics to correct one macro-indicator (ecological<br />

functioning). The here presented evaluation approach provides theoretical background for<br />

further evaluation criteria such as aesthetics, water quality <strong>and</strong> economical wealth. For complex<br />

methods evaluating the functionality of l<strong>and</strong>scapes not only LMs but also further information is<br />

necessary (Lang et al. 2009). Hence, besides a purely quantitative analysis of LMs, scientific<br />

findings such as degrees of hemeroby are essential to make basic assumptions. Employing LMs<br />

without any assumptions <strong>and</strong> restrictions would risk misinterpretations (Fortin et al. 2003). An<br />

additional measure to avoid misinterpretations is the usage of sets of LMs(Lausch <strong>and</strong> Herzog<br />

2002; Cushman et al. 2008). Due to local developed l<strong>and</strong>scape characteristics, employing only<br />

one LM could lead to an over- <strong>and</strong> under-estimation of determined characteristics.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Frank et al. 2010. A regionally adaptable approach of l<strong>and</strong>scape assessment using l<strong>and</strong>scape metrics<br />

39<br />

Table 3: Classification of the human impact on ecosystems <strong>and</strong> the according degree of hemeroby <strong>and</strong><br />

naturalness (Blume <strong>and</strong> Sukopp 1976 -modified)<br />

Hemeroby/ degree of<br />

naturalness<br />

Human impact CORINE- L<strong>and</strong>cover 2000<br />

Ahemerobe/ natural None -<br />

Oligohemerobe/ close<br />

to natural<br />

Mesohemerobe/ seminatural<br />

Very sparsely populated<br />

areas<br />

Sparsely populated<br />

cultural l<strong>and</strong>scapes<br />

e.g. Moors <strong>and</strong> heathl<strong>and</strong>,<br />

Peat bogs<br />

e.g. Complex cultivation patterns,<br />

Natural grassl<strong>and</strong>s<br />

"natural"<br />

Euhemerobe/<br />

far from natural<br />

Polyhemerobe/<br />

strange to nature<br />

Metahemerobe/<br />

artificial<br />

Agricultural l<strong>and</strong>scapes,<br />

settlements<br />

Partly built-up areas,<br />

dump sites<br />

Biocenosis widely<br />

destroyed, inner-cities,<br />

industrial facilities<br />

e.g. Non-irrigated arable l<strong>and</strong>,<br />

Vineyards<br />

e.g. Discontinuous urban fabric,<br />

Dump sites<br />

e.g. Industrial or commercial units,<br />

Road <strong>and</strong> rail networks <strong>and</strong><br />

associated l<strong>and</strong><br />

"not natural"<br />

Figure 1: Hierarchical assessment scheme for the ecosystem service “ecological functioning”<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Frank et al. 2010. A regionally adaptable approach of l<strong>and</strong>scape assessment using l<strong>and</strong>scape metrics<br />

40<br />

Figure 2: Exemplary application of ecological linking matrices for assessing the intensity of l<strong>and</strong> use<br />

Table 4: Evaluation of the habitat connectivity<br />

Habitat connectivity [%] Evaluation<br />

0 - 5 -10<br />

5 - 10 -5<br />

10 – 15 0<br />

15 – 20 5<br />

> 20 10<br />

References<br />

Adger, W.N., Agrawala, S., Mirza, M.M.Q., Conde, C., O'Brien, K.L., Pulhin, J., Pulwarty, R.,<br />

Smit, B., <strong>and</strong> Takahashi, K. 2007. Assessment of adaptation practices, options,<br />

constraints <strong>and</strong> capacity. In Climate <strong>Change</strong> 2007.: Impacts, Adaptation <strong>and</strong><br />

Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the<br />

Intergovernmental Panel on Climate <strong>Change</strong>. (eds. M.L. Parry, O.F. Canziani, J.P.<br />

Palutikof, C.E. Hanson, <strong>and</strong> P.J. van der Linden), pp. 719-743. Cambridge University<br />

Press, Cambridge.<br />

Allen, C.D., Macalady, A.K., Chenchouni, H., Bachelet, D., McDowell, N., Vennetier, M.,<br />

Kitzberger, T., Rigling, A., Breshears, D.D., Hogg, E.H., et al. 2010. A global overview<br />

of drought <strong>and</strong> heat-induced tree mortality reveals emerging climate change risks for<br />

forests. <strong>Forest</strong> Ecology <strong>and</strong> Management 259: 660-684.<br />

Bastian, O., <strong>and</strong> Schreiber, K.-F. 1999. Analyse und ökologische Bewertung der L<strong>and</strong>schaft.<br />

Spektrum Akad. Verl., Heidelberg; Berlin, pp. 564.<br />

Blume, H.P., <strong>and</strong> Sukopp, H. 1976. Ökologische Bedeutung anthropogener<br />

Bodenveränderungen. Schr. Reihe Vegetationskunde 10: 74-89.<br />

BMU. 2007. National Strategy on Biological Diversity. adopted by the federal cabinet on 7<br />

November 2007. Federal Ministry for the Environment, Nature Conversation <strong>and</strong><br />

Nuclear Safety, Bonn.<br />

BNatSchG. 2010. Bundesnaturschutzgesetz. Gesetz über Naturschutz und L<strong>and</strong>schaftspflege.<br />

first version 14.05.1967, actualized 01.03.2010, Bundesgesetzblatt 1/2010, pp. 2542.<br />

Cushman, S.A., McGarigal, K., <strong>and</strong> Neel, M.C. 2008. Parsimony in l<strong>and</strong>scape metrics: Strength,<br />

universality, <strong>and</strong> consistency. Ecological Indicators 8: 691-703.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Frank et al. 2010. A regionally adaptable approach of l<strong>and</strong>scape assessment using l<strong>and</strong>scape metrics<br />

41<br />

Fortin, M.J., Boots, B., Csillag, F., <strong>and</strong> Remmel, T.K. 2003. On the role of spatial stochastic<br />

models in underst<strong>and</strong>ing l<strong>and</strong>scape indices in ecology. Oikos 102: 203-212.<br />

Fuerst, C., Davidsson, C., Pietzsch, K., Abiy, M., Volk, M., Lorz, C., <strong>and</strong> Makeschin, F. 2008.<br />

"Pimp your l<strong>and</strong>scape" - an interactive l<strong>and</strong>-use planning support tool. In Geo-<br />

Environment <strong>and</strong> L<strong>and</strong>scape Evolution III, pp. 219-232. Wit Press.<br />

Goria, A., <strong>and</strong> Gambarelli, G. 2004. Economic Evaluation of Climate <strong>Change</strong> Impacts <strong>and</strong><br />

Adaptation in Italy. FEEM Working Paper No. 103.04. Available at SSRN:<br />

http://ssrn.com/abstract=569122.<br />

Krüsemann, E. 2006. Der Biotopverbund nach § 3 BNatSchG. Natur und Recht 28: 546-554.<br />

Lang, S., Walz, U., Klug, H., Blaschke, T., <strong>and</strong> Syrbe, R.-U. 2009. L<strong>and</strong>scape Metrics - A<br />

toolbox for assessing past, present <strong>and</strong> future l<strong>and</strong>scape structures. In Geoinformation<br />

Technologies for Geocultural <strong>L<strong>and</strong>scapes</strong>: European Perspectives. (eds. O. Bender, N.<br />

Evelpidou, A. Krek, <strong>and</strong> A. Vassilopoulos), pp. 207-234. CRC Press Leiden, Leiden.<br />

Lausch, A., <strong>and</strong> Herzog, F. 2002. Applicability of l<strong>and</strong>scape metrics for the monitoring of<br />

l<strong>and</strong>scape change: issues of scale, resolution <strong>and</strong> interpretability. Ecological Indicators<br />

2: 3-15.<br />

McGarigal, K., <strong>and</strong> Marks, B.J. 1995. FRAGSTATS: spatial pattern analysis program for<br />

quantifying l<strong>and</strong>scape structure. U.S. Department of Agriculture, <strong>Forest</strong> Service,<br />

Pacific Northwest Research Station, Portl<strong>and</strong>, pp. 122.<br />

Molden, D. 2007. Water for food, water for life: A Comprehensive Assessment of Water<br />

Management in Agriculture. Earthscan, London, UK, pp. 645.<br />

Rosenberg, N.J. 1992. Adaptation of agriculture to climate change. Climatic <strong>Change</strong> 21: 385-<br />

405.<br />

SRU. 2000. Umweltgutachten 2000 - Schritte ins nächste Jahrtausend. Metzler-Poeschel,<br />

Stuttgart, pp. 688.<br />

Steinhardt, U., Herzog, F., Lausch, A., Mueller, E., <strong>and</strong> Lehmann, S. 1999. The Hemeroby<br />

Index for L<strong>and</strong>scape Monitoring <strong>and</strong> Evaluation. In Environmental Indices. Systems<br />

Analysis Approach. (eds. Y.A. Pykh, E. Hyatt, <strong>and</strong> R.J.M. Lenz), pp. 237-257. EOLSS<br />

Oxford, Oxford.<br />

Van den Berg, R.D., <strong>and</strong> Feinstein, O. 2009. Evaluating Climate <strong>Change</strong> <strong>and</strong> Development.<br />

Transaction Publications, Piscataway, New Jersey.<br />

Zebisch, M., Wechsung, F., <strong>and</strong> Kenneweg, H. 2004. L<strong>and</strong>scape response functions for<br />

biodiversity--assessing the impact of l<strong>and</strong>-use changes at the county level. L<strong>and</strong>scape<br />

<strong>and</strong> Urban Planning 67: 157-172.<br />

Zomer, R.J., Trabucco, A., Bossio, D.A., <strong>and</strong> Verchot, L.V. 2008. Climate change mitigation: A<br />

spatial analysis of global l<strong>and</strong> suitability for clean development mechanism<br />

afforestation <strong>and</strong> reforestation. Agriculture, Ecosystems & Environment 126: 67-80.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. García et al. 2010. The effect of l<strong>and</strong> cover changes on the habitat of Baird’s tapir in Laguna Lachuá<br />

42<br />

The effect of l<strong>and</strong> cover changes (1960’s-2003) on the habitat<br />

morphological spatial pattern <strong>and</strong> population viability of Baird’s tapir<br />

in Laguna Lachuá National Park Influence Zone, Guatemala<br />

Manolo García 1* , Fern<strong>and</strong>o Castillo 1 , Raquel Leonardo 1 , Liza García 1 & Ivonne<br />

Gómez 1<br />

1 Centro de Datos para la Conservación, Centro de Estudios Conservacionistas,<br />

Universidad de San Carlos, Guatemala<br />

Abstract<br />

The current anthropogenic morphological spatial patterns (MSP) of forests usually are not<br />

suitable for large mammals such as Baird’s tapir, the largest terrestrial mammal in the<br />

Neotropics. We intended to evaluate how changes in MSP affected the population viability (PV)<br />

of this species. Using l<strong>and</strong> cover images we determined the MSP in different years using<br />

software GUIDOS. PV was modeled using software VORTEX. The species habitat reduced<br />

from approximately 95% to 46% of the study area. From 1960 to 1983 the majority of the area<br />

was core with a few perforations. In 1991 more than the 50% of the core was loss, since then,<br />

the reduction of corridors, islets <strong>and</strong> branches, increased the isolation of the park. The PV<br />

changed from 0% to 80% probability of extinction. Results show that habitat modification has<br />

accelerated the extinction process. Redirect this tendency is a challenge to l<strong>and</strong> use planning <strong>and</strong><br />

wildlife conservation.<br />

Keywords: spatial pattern population viability Tapir<br />

1. Introduction<br />

Human activities have modified morphological spatial patterns of natural ecosystems causing<br />

great impacts in wildlife (Cuarón, 2000, McCullough, 1996, Escamilla et al., 2000). One of the<br />

most common disturbances in tropical lowl<strong>and</strong>s at northeastern Guatemala is the transformation<br />

of tropical forest for livestock <strong>and</strong> currently Oil palm plantations, causing habitat reduction <strong>and</strong><br />

fragmentation, as well as its degradation. Large mammals are especially susceptible to habitat<br />

reduction <strong>and</strong> fragmentation (Kinnaird et al., 2003). Baird’s tapir is the largest native terrestrial<br />

mammal in the Neotropics, including Guatemala, with a body size of 2 meter long, 1.5 meters<br />

height <strong>and</strong> a body weight of 100 to 350 kilograms (Emmons, 1990). We intended to evaluate<br />

how changes in l<strong>and</strong> cover from 1960s to 2000s have affected the population viability of this<br />

species.<br />

2. Methodology<br />

2.1 Study area<br />

The study area was the Laguna Lachuá National Park’s (LLNP) influence zone. The LLNP is a<br />

14.5 square kilometers protected area in the north of Guatemala surrounded by rural<br />

communities, mainly Mayan Kekchí, <strong>and</strong> also big farms. It is traversed by a road known as<br />

* Corresponding author. Tel.:+502 55538913<br />

Email address: garcia.manolo@usac.edu.gt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. García et al. 2010. The effect of l<strong>and</strong> cover changes on the habitat of Baird’s tapir in Laguna Lachuá<br />

43<br />

“Franja Transversal del Norte” (Northern Transversal Strip) an area of rural development driven<br />

by the Government in the 60s <strong>and</strong> renovated at present by Oil palm plantations.<br />

2.2 Habitat <strong>and</strong> Population Viability analyses for Baird’s tapir<br />

We obtained digital data of the l<strong>and</strong> cover for the years 1960s, 1983, 1991, 2001 <strong>and</strong> 2003. The<br />

1960s map was digitalized by García et al., (2009) from a printed map, the 1991 <strong>and</strong> 2001<br />

layers were obtained at the <strong>Forest</strong>s National Institute (INAB) GIS database <strong>and</strong> the 2003 from<br />

the Ministery of agriculture, livestock <strong>and</strong> food (MAGA, 2006). All maps were transformed to<br />

raster format with a 500 square meters pixel resolution.<br />

2.2.1 Morphological Spatial Pattern Analysis (MSPA)<br />

We analyzed the Baird’s tapir habitat in the study area according to García et al., (2009), using<br />

the sofware GUIDOS (Vogt et al., 2007; Soille <strong>and</strong> Vogt, 2009). We delimited the minimum<br />

square that includes the LLNP <strong>and</strong> the nearest protected areas which are currently the forest<br />

remnants. We used this quadrangle as study area for MSPA. The MSP for each year was<br />

determined.<br />

2.2.2 Population Viability Analysis<br />

The population viability (PV) was modeled using the software VORTEX (Lacy, Borbat, <strong>and</strong><br />

Pollak, 2005). VORTEX is an individual-based simulation model for PV analysis. The program<br />

requires information about the species reproductive system <strong>and</strong> rates (Miller <strong>and</strong> Lacy, 2005),<br />

we used the data generated during the PV Analysis workshop for Baird’s tapir organized by the<br />

Tapir specialist group held in Belize 2005, <strong>and</strong> modified it with the advice of Arnaud Desbiez<br />

<strong>and</strong> Patricia Medici from the Conservation Breeding Specialist Group from the IUCN. For each<br />

year, we determined the area of the forest remnant that includes the LLNP, <strong>and</strong> estimated the<br />

population of tapirs using the parameters used by García et al., 2010. Threats as hunting <strong>and</strong><br />

carrying capacity reduction were not included to evaluate the effect of the habitat MSP. A<br />

model for each year was run using a time line of 100 years <strong>and</strong> 100 interactions.<br />

3. Result<br />

From the MSPA, changes are dramatic, especially from 1983 to 1991 (see Table 1 <strong>and</strong> Figure 1).<br />

In 1960s the 90% of the area was core forest, in 1983 perforations increased to a 4.5% of the<br />

area, leading to major change to 1991 with only 16.6% of the area covered by core forest.<br />

During this period of time most of the non-core remaining forest was changed to structures such<br />

as bridges. In 1991 there are almost no perforations left, <strong>and</strong> the edge was increased. From<br />

1991 to 2003 the core forest remained almost with the same area, but bridges <strong>and</strong> branches were<br />

constantly declining. Islets at first increased in number but then also decreased which indicates<br />

the total destruction of forest remnants outside the protected area.<br />

For the PV Analysis, changes are related to MSP, which changed dramatically since 1983 (see<br />

Figure 2). From 1960 to 1983 due all the area was suitable habitat, is expected that existed a<br />

continuous population of tapirs including areas of Chiapas in México <strong>and</strong> southern Petén<br />

through Sayaxche <strong>and</strong> La Pasión river in Guatemala. This population exceeded the 1,000<br />

individuals, <strong>and</strong> the resulting model showed to be <strong>and</strong> viable population with 0% probability of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. García et al. 2010. The effect of l<strong>and</strong> cover changes on the habitat of Baird’s tapir in Laguna Lachuá<br />

44<br />

extinction. In 1991 the PV dropped from a initial population of almost 50 individuals with 95%<br />

probability of extinction. At 2003, is estimated to have a population of 15 individuals with a<br />

high probability of extinction over the next 100 years.<br />

4. Discussion<br />

There exists a direct influence of MSP on PV due it determines the population size; both are<br />

useful tools for l<strong>and</strong>scape planning (Beissinger, Nicholson, <strong>and</strong> Possingham, 2009). The results<br />

may show how MSP modifications that occurred in the study area accelerated the natural<br />

extinction process of this species. In the 1960s the population viability had a 100% probability<br />

of surviving in the next 100 years, <strong>and</strong> this was changed when the habitat was transformed.<br />

Currently the remaining population is too small to be viable in the next 100 years, this shows<br />

that the protected area is too small to retain a Tapirs viable population. The fact that the<br />

protected area is too small is a mayor challenge to conservation managers because the influence<br />

zone must be also managed <strong>and</strong> re-ordered in order to increase the PV of large species as<br />

Baird’s tapir. Current pressures such as forest reduction by livestock <strong>and</strong> Oil palm plantations<br />

increases this challenge.<br />

Management should be directed in reverting the l<strong>and</strong> cover change process, increasing core<br />

forest areas <strong>and</strong> connectivity. Species such as Baird’s tapir may be used to design habitat<br />

restoration plans that conduce to future MSP that increases the species PV.<br />

References<br />

Beissinger, S., Nicholson, E., <strong>and</strong> Possingham, H., 2009. Application of population viability<br />

analysis to l<strong>and</strong>scape. In J. Millspaugh, <strong>and</strong> F. Thompson, Models for planning wildllife<br />

conservation in large l<strong>and</strong>scapes (pág. 688).<br />

Cuarón, A. , 2000. Effects of l<strong>and</strong>-cover changes on mammals in neotropical region: a modeling<br />

approach. Conservation Biology , 14 (4), 1676-1692.<br />

Emmons, L., 1990. Neotropical Rainforest Mammals. EEUU: The University of Chicago Press.<br />

Escamilla, A., Sanvicente, M., Sosa, M., <strong>and</strong> Galindo-Leal, C., 2000. Habitat mosaic, wildlife<br />

availability, <strong>and</strong> hunting in tropical forest of Calakmul, Mexico. Conservation Biology ,<br />

14 (6), 1592-1601.<br />

García, M., Leonardo, R., Castillo, F., Gómez, I., <strong>and</strong> García, L., 2010. El tapir<br />

centroamericano (Tapirus bairdii) como herramienta para el fortalecimiento del SIGAP.<br />

Dirección General de Investigación. Universidad de San Carlos de Guatemala.<br />

Guatemala.<br />

García, M., Leonardo, R., Gómez, I., <strong>and</strong> García, L., 2009. Estado actual de conservación del<br />

Tapir (Tapirus bairdii) en el Sistema Guatemalteco de Áreas Protegidas. Fondo<br />

Nacional para la Conservación de la Naturaleza. Guatemala.<br />

Kinnaird, M., S<strong>and</strong>erson, E., O'Brien, T., Wibisono, H., <strong>and</strong> Wollmer, G., 2003. Deforestation<br />

trends in a tropical lanscape <strong>and</strong> implications for endangered large mammals.<br />

Conservation Biology , 17 (1), 245-257.<br />

Lacy, R., Borbat, M., <strong>and</strong> Pollak, J., 2005. VORTEX: A stochastic Simulation of the Extinction<br />

Process. Version 9.5. Brookfield, IL: Chicago Zoological Society.<br />

MAGA , 2006. Mapa de cobertura y Uso de la Tierra de Guatemala 1:50,000. Ministerio de<br />

Agricultura, Ganadería y Alimentación. Gobierno de Guatemala. Guatemala.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. García et al. 2010. The effect of l<strong>and</strong> cover changes on the habitat of Baird’s tapir in Laguna Lachuá<br />

45<br />

McCullough., 1996. Metapopulations <strong>and</strong> wildlife conservation. Isl<strong>and</strong> Press.<br />

Miller, P., <strong>and</strong> Lacy, R., 2005. VORTEX: A stochastic Simulation od the extinction process.<br />

Version User's manual. Apple Valley, MN: Conservation Breeding Specialist Group<br />

(SSC/IUCN).<br />

Soille, P., <strong>and</strong> Vogt, P., 2009. Morphological segmentation of binary patterns. Pattern<br />

recognition letters , 30, 456-459.<br />

Vogt, J., P, Iwanowski, M., Estreguil, C., Kozak, J., <strong>and</strong> Soille, P., 2007. Mapping l<strong>and</strong>scape<br />

corridors. Ecological indicators , 7 (2), 481-488.<br />

Table 1: MSPA parameters <strong>and</strong> their proportion in the study area. Source: Authors<br />

MSPA<br />

% of extencion of study area<br />

parameter 1960s 1983 1991 2001 2003<br />

Core 90.39 86.68 16.58 12.07 13.57<br />

Islet 0 0.01 5.82 6.36 4.09<br />

Perforation 2.57 4.54 0.62 0.34 0.21<br />

Edge 1.61 1.56 3.99 3.36 4.22<br />

Loop 0.4 0.73 1.5 1.35 2.35<br />

Bridge 0.31 0.53 33.96 31.9 18.34<br />

Branch 0.18 0.78 5.05 4.96 3.5<br />

Background 4.54 5.17 32.49 39.67 53.71<br />

Total 100.00 100.00 100.00 100.00 100.00<br />

Figure 1: Morphological Spatial Pattern Analysis parameters <strong>and</strong> their proportion in the study area from<br />

1960s to 2003. Source: Authors<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. García et al. 2010. The effect of l<strong>and</strong> cover changes on the habitat of Baird’s tapir in Laguna Lachuá<br />

46<br />

Figure 2: Population viability showing the probability of surviving over the next 100 years for different<br />

models corresponding to the years 1960s, 1983, 1991, 2001 <strong>and</strong> 2003. Source: Authors<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Gholami et al. 2010. Spatial pattern of soil macrofauna biodiversity in wildlife refugee of Karkhe in Southwestern Iran<br />

47<br />

Spatial pattern of soil macrofauna biodiversity in wildlife refugee of<br />

Karkhe in Southwestern Iran<br />

Shaieste Gholami 1* , Seied Mohsen Hosseini 1 , Jahangard Mohammadi 2 & Abdolrassoul<br />

Salman Mahini 3<br />

1 Tarbiat Modares University, Iran<br />

2 Shahrekord University, Iran<br />

3 Gorgan University, Iran<br />

Abstract<br />

Information about the spatial patterns of soil biodiversity is limited though required, e.g. for<br />

underst<strong>and</strong>ing effects of biodiversity on ecosystem processes. This study was conducted to<br />

determine whether soil macrofauna biodiversity parameters display spatial patterns in the<br />

riparian forest l<strong>and</strong>scape of Karkhe, southwestern Iran. Soil macrofauna were sampled in 2009<br />

using 200 sampling point along parallel transects (perpendicular to the river). Maximum<br />

distance between samples was 0.5 km. Soil macrofauna were extracted from 50 cm×50 cm×25<br />

cm soil monolith by h<strong>and</strong>-sorting procedure. Abundance <strong>and</strong> Shannon H’ index were analyzed<br />

using geostatistics (variogram) in order to describe <strong>and</strong> quantify the spatial continuity. The<br />

variograms were spherical <strong>and</strong> revealed the presence of spatial autocorrelation. The range of<br />

influence was 1724 m for abundance <strong>and</strong> 1326 m for diversity. The variograms featured high<br />

ratio of nugget variance to sill. This showed that there was the small-scale variability <strong>and</strong><br />

proportion of unexplained variance.<br />

Keywords: Biodiversity, Soil macrofauna, Spatial pattern, Variogram<br />

1. Introduction<br />

In the last 15-20 years, riparian forests have become recognized as important components of<br />

l<strong>and</strong>scape <strong>and</strong> serve as a vital link between the aquatic environment <strong>and</strong> upl<strong>and</strong> ecosystems<br />

(Giese et al. 2000). Riparian ecosystems are aquatic-terrestrial ecotones with unique biotic,<br />

biophysical <strong>and</strong> l<strong>and</strong>scape characteristics (Lyon <strong>and</strong> Gross 2005). Accordingly, sustainability<br />

<strong>and</strong> maintenance of riparian vegetation or restoring of degraded sites is critical to sustain<br />

inherent ecosystem function <strong>and</strong> values (Giese et al. 2000). Many scientists believe promoting<br />

sustainability is the overarching goal of l<strong>and</strong>scape (<strong>and</strong> regional) planning, (Leitao <strong>and</strong> Ahern<br />

2002).<br />

A current challenge in ecology is underestimating how patterns <strong>and</strong> processes vary with scale<br />

<strong>and</strong> searching for general principles (assembly rules) which determine the species composition<br />

of communities (Guillaume 2002). Also a fundamental <strong>and</strong> unanswered research question in<br />

l<strong>and</strong>scape ecology centers on how the spatial arrangement of the ecosystems influences the<br />

distribution <strong>and</strong> abundance of organisms (Coulson et al. 1999). Description of patterns in<br />

species assemblages <strong>and</strong> diversity is an essential step before generating hypotheses in functional<br />

ecology (Guillaume 2002).<br />

If we want to have information about ecosystem function, soil biodiversity is best considered by<br />

focusing on the groups of soil organisms that play major roles in ecosystem functioning when<br />

exploring links with provision of ecosystem services (Barrios 2007). Information about the<br />

* Corresponding author. Tel.:+989163716945 - Fax:+986712231662<br />

Email address: Mozhgangholami@yahoo.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Gholami et al. 2010. Spatial pattern of soil macrofauna biodiversity in wildlife refugee of Karkhe in Southwestern Iran<br />

48<br />

spatial pattern of soil biodiversity at the regional scale is limited though required, e.g. for<br />

underst<strong>and</strong>ing regional scale effects of biodiversity on ecosystem processes (Joschko et.al.<br />

2006). The practical consequences of these findings are useful for sustainable management of<br />

soils <strong>and</strong> in monitoring soil quality. Soil macrofauna play significant, but largely ignored, roles<br />

in the delivery of ecosystem services by soils at plot <strong>and</strong> l<strong>and</strong>scape Scales (Lavelle et.al. 2006).<br />

One main reason responsible for the absence of information about biodiversity at regional scale,<br />

is the lack of adequate methods for sampling <strong>and</strong> analyzing data at this dimension. An adequate<br />

approach for the analysis of spatial patterns is a transect study in which samples are taken in a<br />

certain order <strong>and</strong> with a certain distance between samples (Joschko et.al. 2006). Geostatistics<br />

provide descriptive tools such as variogram to characterize the spatial pattern of continuous <strong>and</strong><br />

categorical soil attributes (Goovaerts 1999; Gringarten <strong>and</strong> Deutsch 2001; Mohammadi 2006).<br />

This method allows assessment of consistency of spatial patterns as well as the scale at which<br />

they are expressed (Jimenez et al. 2001).<br />

The aim of this study is to analyze spatial patterns of soil macrofauna (= invertebrates visible at<br />

the naked eye) in Wildlife Refugee of Karkhe in the riparian forest of the southwestern Iran.<br />

Parameters of soil macrofauna biodiversity comprise: abundance (total abundance of<br />

macrofauna), <strong>and</strong> diversity.<br />

2. Methodology<br />

The study was carried out in Wildlife Refugee of Karkhe in the riparian forest of the<br />

southwestern Iran (31 o 57 / - 32 o 05 / N <strong>and</strong> 48 o 13 / - 48 o 16 / E). The climate of the study area is<br />

semi-arid. Average yearly rainfall is about 325.5 mm with a mean temperature of 24 oc . Plant<br />

cover, mainly comprises Populus euphratica <strong>and</strong> Tamarix sp.<br />

The both sides of river are similar, so we sampled on one of the two sides. Soil macrofauna<br />

were sampled in 2009 using 200 sampling point along parallel transects (perpendicular to the<br />

river). The distance between transects were 0.5 km. The sampling procedure was hierarchically,<br />

we considered maximum distance between samples as 0.5 km, but the samples was taken at<br />

250m, 100m, 50m, 20m, 15m, 10m, 5m, 2m <strong>and</strong> 1m at different location of sampling.<br />

soil macrofauna (= invertebrates visible at the naked eye) was extracted from 50 cm×50 cm×25<br />

cm soil monolith by h<strong>and</strong>-sorting procedure at the last winter (because at this time moisture <strong>and</strong><br />

temperature are suitable <strong>and</strong> soil macrofauna reach their highest abundance). All soil<br />

macrofauna were identified to family level.<br />

Number of animals (abundance) <strong>and</strong> diversity (Shannon H’ index) by using PAST version 1.39,<br />

were determined in each sample. Classical statistical parameters, i.e. mean, st<strong>and</strong>ard deviation,<br />

coefficient of variation, minimum <strong>and</strong> maximum, were calculated using SPSS17 software.<br />

Diversity <strong>and</strong> abundance data were analyzed using geostatistics (variogram) in order to describe<br />

<strong>and</strong> quantify the spatial continuity. Geostatistical analysis was performed using the software<br />

Variowin 2.2 (variograms). Spatial distribution maps were made by block kriging using the<br />

software Geoease <strong>and</strong> Surfer 8.0.<br />

3. Result<br />

Soil macrofauna communities were dominated by earthworm, diplopods, coleoptera, gastropoda,<br />

araneae, <strong>and</strong> insect larvae, reaching an abundance of 43.1 individuals / m 2 .<br />

Table 1 shows the mean, st<strong>and</strong>ard deviation, coefficient of variation, minimum <strong>and</strong> maximum<br />

values for soil macrofauna abundance <strong>and</strong> diversity.<br />

The variograms revealed the presence of spatial autocorrelation Fig. 1. The parameters of the<br />

theoretical models fitted to experimental variograms are given in Table 2. The variograms of<br />

two indices were spherical <strong>and</strong> showed positive nugget, which can be explained by sampling<br />

error, short range variability, r<strong>and</strong>om <strong>and</strong> inherent variability. The nugget-to-sill ratio can be<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Gholami et al. 2010. Spatial pattern of soil macrofauna biodiversity in wildlife refugee of Karkhe in Southwestern Iran<br />

49<br />

used to classify the spatial dependence of soil properties. In this study we used similar criteria to<br />

those reported by Sun et al., (2003). The variable is considered to have a strong spatial<br />

dependence if the ratio is less than 25%, <strong>and</strong> has a moderate spatial dependence if the ratio is<br />

between 25% <strong>and</strong> 75%; otherwise, the variable has a weak spatial dependence.<br />

Soil macrofauna abundance <strong>and</strong> diversity were moderately spatially dependent (Table 2). The<br />

range of influence is considered as the distance beyond which observations are not spatially<br />

dependent. This distance ranged from 1326 m for soil macrofauna diversity to 1724 m for<br />

abundance (Table 2).<br />

The maps obtained by kriging for soil macrofauna abundance <strong>and</strong> diversity are shown in Fig. 2.<br />

Table 1. Mean, st<strong>and</strong>ard deviation (S.D.), coefficient of variation (CV), minimum <strong>and</strong> maximum values<br />

of for soil macrofauna abundance <strong>and</strong> diversity<br />

Index mean S. D C.V (%) minimum maximum<br />

Abundance (indiv. m -2 ) 43.1 73.9 171 0 480<br />

Shannon (H’) 0.55 0.51 92 0 1.9<br />

Table2. Variogram model parameters for soil macrofauna abundance <strong>and</strong> diversity<br />

Index Model Sill Nugget Nugget/ Sill (%) Range(m)<br />

abundance spherical 1.13 0.59 52 1724<br />

Shannon (H’) spherical 0.27 0.153 55 1326<br />

(A)<br />

(B)<br />

Fig.1. Variograms of (A) Log transformed data for soil macrofauna abundance <strong>and</strong> (B) soil macrofauna<br />

diversity<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Gholami et al. 2010. Spatial pattern of soil macrofauna biodiversity in wildlife refugee of Karkhe in Southwestern Iran<br />

50<br />

3554000<br />

3554000<br />

3552000<br />

3552000<br />

3550000<br />

3550000<br />

3548000<br />

River<br />

3548000<br />

River<br />

3546000<br />

3546000<br />

3544000<br />

3<br />

3544000<br />

1<br />

2.2<br />

0.8<br />

3542000<br />

3542000<br />

0.6<br />

N<br />

3540000<br />

0 1000 2000<br />

237000 239000 241000 243000<br />

1.4<br />

0.6<br />

N<br />

3540000<br />

0 1000 2000<br />

237000 239000 241000 243000<br />

0.4<br />

0.2<br />

4. Discussion<br />

Fig.2. Distribution of soil macrofauna diversity: A – Abundance, B –Shannon H index<br />

This study showed that soil macrofauna abundance <strong>and</strong> diversity were spatially autocorrelated<br />

within the range of 1519 <strong>and</strong> 1937 meters. In line with our result Gongalski et.al., (2008)<br />

reported the spatial autocorrelation for soil macrofauna abundance <strong>and</strong> diversity in<br />

Mediteranean forest in Russia.<br />

Typically these structures constitute one source of the nugget variance of the variograms (Rossi,<br />

2003). However, the variograms reported here featured a somewhat high ratio of nugget<br />

variance to sill (Table 2). This result showed that there was the small-scale variability <strong>and</strong><br />

important proportion of unexplained variance (Rossi, 2003, Gongalski et.al., 2008,).<br />

Spatial distribution of soil macrofauna at large scales may be influenced by factors like<br />

gradients in soil organic matter (quantity <strong>and</strong> or quality), texture <strong>and</strong> vegetation cover structure<br />

(Ettema & Wardle 2002). These factors together with intrinsic population processes constitute<br />

proximate controlling factors of population structure (Rossi, 2003).<br />

References<br />

Barrios, E. 2007. Soil biota, ecosystem services <strong>and</strong> l<strong>and</strong> productivity. Ecological Economics,<br />

24(2):269-285.<br />

Coulson, R.N., McFadden, B.A., Pully, P.A., Lovelady, C.N., Fitzgerald, J.W., Jack, S.B., 1999.<br />

Heterogeneity of forest l<strong>and</strong>scape <strong>and</strong> the distribution <strong>and</strong> abundance of southern pine<br />

beetle. <strong>Forest</strong> Ecology <strong>and</strong> Management, 114: 471-485.<br />

Ettema, C. H., Wardle, D. A. 2002. Spatial soil ecology. Trends in Ecology <strong>and</strong> Evolution 17,<br />

177–183.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Gholami et al. 2010. Spatial pattern of soil macrofauna biodiversity in wildlife refugee of Karkhe in Southwestern Iran<br />

51<br />

Giese, L.A., Aust, W.M., Trettin, C.C., Kolka, R.K., 2000. Spatial <strong>and</strong> temporal patterns of<br />

carbon storage <strong>and</strong> species richness in three South Carolina coastal plain riparian forests.<br />

Ecological Engineering, 15: S157-S17.<br />

Gonglanski, K.B., Gorshkova, I.A,. Karpov, A.I., Pokarzhevskii, A.D. 2008. Do boundaries of<br />

soil animal <strong>and</strong> plant communities coincide A case study of a Mediterranean forest in<br />

Russia. European j ournal of soil biology.44:355–363.<br />

Goovaerts, P., 1999. Geostatistics in soil science: state-of-the-art <strong>and</strong> perspectives. Geoderma,<br />

89: 1-45.<br />

Gringarten, E. <strong>and</strong> Deutsch, C.V., 2001. Teacher's aide, Variogram interpretation <strong>and</strong> modeling.<br />

Mathematical Geology, 33(4):507-534.<br />

Guillaume, D., 2002. Patterns of plant species <strong>and</strong> community diversity at different organization<br />

levels in a forested riparian l<strong>and</strong>scape. Journal of Vegetation Science, 13: 91-106.<br />

Jimenez, J.J., Rossi, J.P., Lavelle, P., 2001. Spatial distribution of earthworm in acid-soil<br />

savannas of the eastern plains of Colombia. Applied Soil Ecology, 17: 267-278.<br />

Joschko, M., Fox, C.A., Lentzsch, P., Kiesel, J., Hierold, W., Kruck, S., Timmer, j., 2006.<br />

Spatial analysis of earthworm biodiversity at the regional scale. Agriculture Ecosystem &<br />

Environment, 112: 367-380.<br />

Lavelle, P., Decaënsb, T., Aubert, M., Barot, S., Blouin, M., Bureau, F., Margerieb, P.,<br />

Mora, P., Rossi, J.-P 2006. Soil invertebrates <strong>and</strong> ecosystem services. European Journal of Soil<br />

Biology 42: S3–S15.<br />

Leitao, A.B. <strong>and</strong> Ahern, J., 2002. Applying l<strong>and</strong>scape ecological concepts <strong>and</strong> metrics in<br />

sustainable l<strong>and</strong>scape planning. L<strong>and</strong>scape And Urban Planning, 59: 65-93.<br />

Lyon, J. <strong>and</strong> Gross, N.M., 2005. Patterns of plant diversity <strong>and</strong> plant-environmental relationship<br />

across three riparian corridors. <strong>Forest</strong> Ecology <strong>and</strong> Management, 204: 267-278.<br />

Mohmmadi, J. 2006. Pedometrics: Spatial statistics (Geostatistics). Pelk Publications,453p.<br />

Rossi, J.P., 2003. Short-range structures in earthworm spatial distribution. Pedo biologia, 47:<br />

582-587.<br />

Sun, B., Zhou, S., Zhao, Q., 2003. Evaluation of spatial <strong>and</strong> temporal changes of soil quality<br />

based on geostatistical analysis in the hill region of subtropical China. Geoderma, 115:<br />

85-99.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. González-Moreno et al. 2010. The influence of spatial structure on natural regeneration <strong>and</strong> biodiversity<br />

52<br />

The influence of spatial structure on natural regeneration <strong>and</strong><br />

biodiversity in Mediterranean pine plantations: a nested l<strong>and</strong>scape<br />

approach<br />

P. González-Moreno 1,* ; J. L. Quero 2,3 , L. Poorter 2 , F.J. Bonet 1 & R. Zamora 4<br />

1 Centro Andaluz de Medio Ambiente, Granada, Spain<br />

2 <strong>Forest</strong> Ecology <strong>and</strong> <strong>Forest</strong> Management Group, Centre for Ecosystem Studies,<br />

Wageningen University, Wageningen, The Netherl<strong>and</strong>s<br />

3 Área de Biodiversidad y Conservación, Departamento de Biología y Geología,<br />

Universidad Rey Juan Carlos, Móstoles, Spain<br />

4 Departmento de Ecología, Universidad de Granada, Granada, Spain<br />

Abstract<br />

Promoting plant diversity in plantations is a worldwide concern. This research aimed to evaluate<br />

the effect of the spatial configuration of Mediterranean pine plantations on regeneration <strong>and</strong><br />

plant diversity in order to facilitate management decisions. Spatial characteristics of pine<br />

plantation patches at l<strong>and</strong>scape scale (distances to other vegetation types) <strong>and</strong> at patch scale<br />

(patch geometry <strong>and</strong> internal structure) were related to abundance of Quercus ilex seedlings <strong>and</strong><br />

the Shannon diversity index of plant species. Results showed that Q. ilex regeneration <strong>and</strong> plant<br />

diversity are affected by the spatial configuration. (1) Proximity to oak patches favoured<br />

abundance of Q. ilex seedlings <strong>and</strong> plant diversity. (2) Patch geometry affected plant diversity,<br />

with larger patches having less diversity. (3) Internal structure influenced both regeneration of<br />

Q. ilex <strong>and</strong> diversity. More heterogeneous areas were characterized by higher diversity <strong>and</strong> less<br />

abundant oak regeneration suggesting that some species show different response to microhabitat<br />

heterogeneity.<br />

Keywords: fragmentation; texture; context; geometry; spatial configuration<br />

1. Introduction<br />

Pine plantations cover a large extent in the Mediterranean basin, representing ca. 12% of the<br />

total forest cover (FAO 2006). A large extent of these plantations is the result of reforestation<br />

programs, carried out since the 19 th century (Pausas et al. 2004). Current management trends are<br />

based on the multifunctionality of these plantations considering not only the original protective<br />

<strong>and</strong> productive functions but also other factors such as biodiversity or recreation (Brockerhoff et<br />

al. 2008). To facilitate plantation management towards mixed st<strong>and</strong>s it is necessary to<br />

underst<strong>and</strong> which factors affect natural regeneration <strong>and</strong> plant diversity. One of these<br />

influencing factors is the spatial pattern of pine plantations at different scales (Lookingbill <strong>and</strong><br />

Zavala 2000) For instance, pine plantation patches can be found in an heterogeneous mosaic of<br />

different vegetation types that will influence their dynamic <strong>and</strong> ecological characteristics. This<br />

effect at l<strong>and</strong>scape scale could be assessed by two aspects: vegetation context <strong>and</strong> patch<br />

geometry.<br />

Vegetation context is the relation among plantation patches <strong>and</strong> other vegetation types.<br />

Proximity to specific vegetation types will facilitate propagules exchange (White et al. 2004).<br />

* Corresponding author. Tel: +34 958241000 - Fax: +34 958137246<br />

Email address: glezmoreno@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. González-Moreno et al. 2010. The influence of spatial structure on natural regeneration <strong>and</strong> biodiversity<br />

53<br />

However, proximity relations could be complicated considering the relief. Mountain areas are<br />

anisotropic surfaces where the downhill dispersal of propagules will be easier due to the direct<br />

effect of gravity (Ohsawa et al. 2007) or because animal dispersal vectors move downhill in<br />

order to save energy (Li <strong>and</strong> Zhang 2003). Patch geometry considers the shape or area of<br />

patches <strong>and</strong> the effect that those characteristics could have on internal patch dynamics. Patch<br />

geometry determines the edge effect (Turner, Gardner, <strong>and</strong> O'Neill 2001). The edge has drier<br />

conditions, with more light <strong>and</strong> it receives the visit of open habitat species. These differences<br />

influence species composition <strong>and</strong> thus biodiversity <strong>and</strong> probability of propagules arrival.<br />

Turner et al., (2001: 3) defines l<strong>and</strong>scape as an area that is spatially heterogeneous in at least<br />

one factor of interest. This heterogeneity, can be observed at different scales. Thus, given a<br />

l<strong>and</strong>scape with a determined scale, it is possible to identify mosaics of patches within patches.<br />

This nested model of mosaics is important to underst<strong>and</strong> the ecological processes because it<br />

implies that the spatial configuration of mosaics at different scales are interconnected.<br />

Therefore, not only the vegetation context <strong>and</strong> patch geometry will affect the performance of<br />

recruitment of species but also the mosaic of microhabitats within plantation patches (i.e.<br />

internal vegetation structure). This structural diversity can be obtained from texture analysis of<br />

high resolution imagery (Hepinstall <strong>and</strong> Sader 1997; St-Louis et al. 2006). Texture is the spatial<br />

distribution of different gray-levels in the same b<strong>and</strong> of the image (Haralick, Shanmugam, <strong>and</strong><br />

Dinstein 1973). Considering that each gray-level is the spectral response to a specific vegetation<br />

type or microhabitat, the analysis of the spatial combination of gray-levels can give valuable<br />

information about the spatial structure. This type of analysis can be done applying the Gray<br />

Level Co-occurrence Matrix (GLCM) developed by (Haralick, Shanmugam, <strong>and</strong> Dinstein<br />

1973). This method gives the probability that two pixels in the same image window have same<br />

tone in a given distance <strong>and</strong> direction.<br />

In a previous research in the same study area we evaluated the effect of several environmental<br />

gradients (climate, distance to oak vegetation <strong>and</strong> st<strong>and</strong> density) on biodiversity <strong>and</strong> plant<br />

regeneration of pine plantations (Gómez-Aparicio et al., 2009). Here our objective was to<br />

evaluate specifically the effect of the spatial configuration of pine plantations on tree<br />

regeneration <strong>and</strong> plant diversity at different scales. Specifically we asked how is natural<br />

regeneration <strong>and</strong> plant diversity of different species within plantation patches related at<br />

l<strong>and</strong>scape scale to the vegetation context of pine plantation patches (proximity to seed source)<br />

<strong>and</strong> plantation patch geometry (area <strong>and</strong> patch shape complexity) <strong>and</strong> at patch scale to the<br />

internal structural diversity of pine plantations patches<br />

2. Methodology<br />

2.1 Study site<br />

Sierra Nevada National Park (Southeast Spain) is a mountain region with an altitudinal range<br />

between 860 m <strong>and</strong> 3482 m. It has an extension of more than 2000 Km 2 , <strong>and</strong> a main length of<br />

90 Km. Annual average temperature decreases in altitude from 12-16 ºC bellow 1500 m to 0 ºC<br />

above 3000 m. Precipitation is scarce in summer, while the winter precipitation is mainly in<br />

form of snow over 2000 m. The average annual precipitation oscillates from less than 250 mm<br />

in the lowest <strong>and</strong> Eastern part of the mountain, to more than 700 mm in the highest peaks.<br />

2.2 Dataset<br />

Regeneration <strong>and</strong> plant diversity variables were obtained from the <strong>Forest</strong> Inventory of Sierra<br />

Nevada National Park collected during 2004-2005 (SINFONEVADA). 275 inventory plots<br />

within pine plantation were selected for the analysis covering a gradient from 974 to 2439 m<br />

a.s.l. Plot size ranged from 300 to 400 m 2 . Two additional subplots were established within each<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. González-Moreno et al. 2010. The influence of spatial structure on natural regeneration <strong>and</strong> biodiversity<br />

54<br />

larger plot: a 5-m radius circle to measure the number of saplings (DBH = 2.5-7.5 cm) <strong>and</strong><br />

seedlings (DBH < 2.5 <strong>and</strong> height < 1.3 m) of tree species, <strong>and</strong> a 10-m radius plot to measure the<br />

species composition <strong>and</strong> abundance by the Braun-Blanquet cover-abundance scale.<br />

Regeneration within pine plantations was measured as seedling abundance. The species<br />

considered in the analysis was Quercus ilex subsp. ballota (Desf.) Samp. (Q. ilex). Other major<br />

species were not considered due to their low abundance in the inventory. Saplings were not<br />

considered, since oak saplings could have been established before the establishment of the pine<br />

plantation, <strong>and</strong> pine “saplings” could be suppressed old planted individuals. Plant diversity was<br />

measured using the Shannon diversity index for the total of species <strong>and</strong> considering only<br />

herbaceous species, only flesh-fruited woody species or only dry-fruited woody species. The<br />

distinction among woody species was considered in order to account the differences in dispersal<br />

syndrome. Flesh-fruited woody species usually have endozoochorous syndromes whilst rest of<br />

species can have other syndromes such as exozoochorous or anemochorous. This distinction<br />

was not made for herbaceous species because most of them (> 95 %) have dry fruits <strong>and</strong> abiotic<br />

dispersal.<br />

A simplification of the forest vegetation map of Andalusia 1:10 000 (CMA 2001) was used to<br />

identify pine plantation areas <strong>and</strong> calculate patch geometry <strong>and</strong> vegetation context variables.<br />

The different vegetation classes used were selected according to their possible contribution to<br />

regeneration <strong>and</strong> plant diversity in pine plantations. The vegetation classes considered were: (1)<br />

pine plantation (> 50 % tree cover <strong>and</strong> > 75% conifers), (2) Oak <strong>and</strong> broadleaved species (> 5 %<br />

tree cover <strong>and</strong> oak presence), (3) shrubl<strong>and</strong>s (< 5 % tree cover, > 20 % shrub cover <strong>and</strong>


P. González-Moreno et al. 2010. The influence of spatial structure on natural regeneration <strong>and</strong> biodiversity<br />

55<br />

diversity variables using correlation analyses. Regeneration <strong>and</strong> plant diversity were calculated<br />

as the average of the values of all inventory plots within each patch.<br />

3. Results<br />

3.1 Vegetation context<br />

Plant diversity declined with increasing distances to oak vegetation, riparian vegetation <strong>and</strong><br />

shrubl<strong>and</strong> (Table 1). However, proximity to riparian vegetation was not influencing herbaceous<br />

diversity. Surprisingly, the diversity index for flesh-fruited woody species had significant<br />

positive relationship to distance to agriculture fields. Considering distance to oak, shrubl<strong>and</strong> <strong>and</strong><br />

riparian vegetation, the algorithm favouring downhill dispersion had higher correlation strength<br />

than the others (weighted downhill > Euclidean > weighted uphill). In contrast to plant diversity<br />

indices, seedling abundance of Q. ilex showed only a strong negative relationship to distances to<br />

oak (Table 1). Abundance of Q. ilex seedlings showed also similar pattern of differences among<br />

algorithms than plant diversity indices (Euclidean = weighted downhill < weighted uphill).<br />

3.2 Patch geometry<br />

Patch area showed negative relationship with Shannon diversity index for all species (ρ=-0.59,<br />

P=0.05), herbaceous species (ρ=-0.60, P=0.05), <strong>and</strong> woody species although the latter<br />

relationship was not significant. Shape index did not correlate significantly with any of the<br />

Shannon diversity indices. Abundance of Q. ilex seedlings did not present significant relation<br />

with any of the geometry variables considered, although there was a positive trend with shape<br />

index (ρ=0.13, P


P. González-Moreno et al. 2010. The influence of spatial structure on natural regeneration <strong>and</strong> biodiversity<br />

56<br />

Euclidean) suggest that seed dispersal is favoured from species-rich patches occurring at higher<br />

altitude towards lower situated pine plantations. Secondly, patch geometry might affect seed<br />

permeability <strong>and</strong> therefore the overall regeneration dynamic of the patch. According to our<br />

results, fragmentation of pine plantations (i.e. patch area reduction) increases overall plant<br />

diversity. Thus, increasing edge effects in pine plantations will facilitate higher rates of plant<br />

diversity. Patch geometry effects on natural processes are based on the delimitation of isolated<br />

discrete units or patches. However some types of l<strong>and</strong>scapes do not present clear patch<br />

delimitation (Gustafson 1998). In our study site, l<strong>and</strong>scape is better depicted as a continuum of<br />

patches with different perturbation rate <strong>and</strong> following a belt structure. This limitation was<br />

overcome selecting isolated patches across the study site. This approach allowed the study of<br />

the effect of geometry on regeneration <strong>and</strong> plant diversity but also pointed out the complexity of<br />

vegetation pattern at l<strong>and</strong>scape scale that eventually might be better depicted as a continuous<br />

gradient of point-data (Gustafson 1998). Thirdly, once propagules are trapped in pine plantation<br />

patches, seeds will require special conditions to germinate <strong>and</strong> establish. This study has proven<br />

that internal vegetation structure measured in terms of texture indices, might be useful to<br />

estimate regeneration <strong>and</strong> plant diversity. All plant diversity indices but for fleshly-fruited<br />

woody species were higher in plots of higher heterogeneity. Areas with higher microhabitat<br />

diversity might have a higher abundance of niches for different species that in turns will<br />

influence positively plant diversity. This finding agrees with the extensive literature that points<br />

the positive relationship between habitat heterogeneity <strong>and</strong> species abundance <strong>and</strong> distribution<br />

(Noss 1990). Nevertheless, this effect proved to be species-dependent. Q. ilex responded in an<br />

opposite manner with higher regeneration rates in structurally homogeneous plantation patches.<br />

Despite these promising findings <strong>and</strong> the theoretical usefulness of texture indices as<br />

heterogeneity quantification techniques (Turner 1991), further research is needed to test this<br />

methodology in other l<strong>and</strong>scapes <strong>and</strong> at different scales to firmly confirm their reliability.<br />

Table 1. Pearson correlation coefficients between regeneration (Log seedling abundance of Q. ilex) <strong>and</strong><br />

biodiversity (Shannon diversity index for all species, herbaceous, dry-fruited <strong>and</strong> flesh-fruited woody<br />

species) variables <strong>and</strong> vegetation context (distances) <strong>and</strong> internal structure variables (entropy <strong>and</strong><br />

contrast). Riparian distance was square root transformed <strong>and</strong> rest of context variables were double square<br />

root transformed. n=275. (*P


P. González-Moreno et al. 2010. The influence of spatial structure on natural regeneration <strong>and</strong> biodiversity<br />

57<br />

Acknowledgements<br />

This research work has been done in the framework of GESBOME Project (RNM 1890) from<br />

the Excellence Research Group Programme of the Andalusian Government. Financial support<br />

was provided by Caja Madrid foundation to PGM <strong>and</strong> by postdoc contract (2007-0572) from the<br />

Spanish Ministry of Science <strong>and</strong> Innovation to JLQ. We are also very grateful to TRAGSA for<br />

conducting the <strong>Forest</strong> Inventory <strong>and</strong> to Lorena Gómez Aparicio for her advices.<br />

References<br />

Brockerhoff, E. et al., 2008. Plantation forests <strong>and</strong> biodiversity: oxymoron or opportunity.<br />

Biodiversity <strong>and</strong> Conservation 17: 925-951.<br />

FAO, 2006. <strong>Global</strong> <strong>Forest</strong> Resources Assessment 2005- Progresss towards sustainable forest<br />

management. FAO. Rome, Italy.<br />

Gómez-Aparicio, L. et al., 2009. Are pine plantations valid tools for restoring Mediterranean<br />

forests An assessment along abiotic <strong>and</strong> biotic gradients. Ecological Applications 19:<br />

2124-2141.<br />

Gustafson, E. J., 1998. Quantifying L<strong>and</strong>scape Spatial Pattern: What Is the State of the Art.<br />

Ecosystems 1: 143-156.<br />

Haralick, R. M., Shanmugam, K. <strong>and</strong> Dinstein, I., 1973. Textural features for image<br />

classification. IEEE Transactions on systems, man, <strong>and</strong> cybernetics smc-3: 610-621.<br />

Hepinstall, A., <strong>and</strong> Sader, S. A., 1997. Using Bayesian statistics, thematic mapper satellite<br />

imagery, <strong>and</strong> breeding bird survey data to model bird species probability of occurrence<br />

in Maine. Photogrammetric Engineering & Remote Sensing 63: 1231-1237.<br />

Hewitt, N. <strong>and</strong> Kellman, M., 2002. Tree seed dispersal among forest fragments: II. Dispersal<br />

abilities <strong>and</strong> biogeographical controls. Journal of Biogeography 29: 351-363.<br />

Hill, J. L. <strong>and</strong> Curran, P. J., 2003. Area, shape <strong>and</strong> isolation of tropical forest fragments: effects<br />

on tree species diversity <strong>and</strong> implications for conservation. Journal of Biogeography 30:<br />

1391-1403.<br />

Li, H., <strong>and</strong> Zhang, Z., 2003. Effect of rodents on acorn dispersal <strong>and</strong> survival of the Liaodong<br />

oak (Quercus liaotungensis Koidz.). <strong>Forest</strong> Ecology <strong>and</strong> Management 176: 387-396.<br />

Lookingbill, T. R., <strong>and</strong> Zavala, M.A., 2000. Spatial Pattern of Quercus ilex <strong>and</strong> Quercus<br />

pubescens Recruitment in Pinus halepensis Dominated Woodl<strong>and</strong>s. Journal of<br />

Vegetation Science 11: 607-612.<br />

Noss, R. F., 1990. Indicators for Monitoring Biodiversity: A Hierarchical Approach.<br />

Conservation Biology 4: 355-364.<br />

Ohsawa, T. et al., 2007. Steep slopes promote downhill dispersal of Quercus crispula seeds <strong>and</strong><br />

weaken the fine-scale genetic structure of seedling populations. Annals of <strong>Forest</strong><br />

Science 64: 405.<br />

Pausas, J. et al., 2004.Pines <strong>and</strong> oaks in the restoration of Mediterranean l<strong>and</strong>scapes of Spain:<br />

New perspectives for an old practice — a review. Plant Ecology 171: 209-220.<br />

St-Louis, V. et al., 2006. High-resolution image texture as a predictor of bird species richness.<br />

Remote Sensing of Environment 105: 299-312.<br />

Turner, M. G. 1991., Quantitative methods in l<strong>and</strong>scape ecology. Springer.<br />

Turner, M. G., Gardner, R. H. <strong>and</strong> O'Neill, R. V., 2001. L<strong>and</strong>scape Ecology in Theory <strong>and</strong><br />

Practice: Pattern <strong>and</strong> Process. New York: Springer-Verlag NY.<br />

White, W. et al., 2004. Seed dispersal to revegetated isolated rainforest patches in North<br />

Queensl<strong>and</strong>. <strong>Forest</strong> Ecology <strong>and</strong> Management 192: 409-426<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Höbinger et al. 2010. Impact of changing cultivation systems on the l<strong>and</strong>scape structure of La Gamba<br />

58<br />

Impact of changing cultivation systems on the l<strong>and</strong>scape structure of<br />

La Gamba, southern Costa Rica<br />

Tamara Höbinger 1* , Stefan Schindler 1 & Anton Weissenhofer 2,3<br />

1 Department of Conservation Biology, Vegetation & L<strong>and</strong>scape Ecology, Faculty of<br />

Life Sciences, University of Vienna, Rennweg 14, A-1030 Vienna, Austria<br />

2 Department of Structural <strong>and</strong> Functional Botany, Faculty of Life Sciences, University<br />

of Vienna, Rennweg 14, A-1030 Vienna, Austria<br />

3 Tropical Station La Gamba, Postal 178, Golfito/Puntarenas, Costa Rica<br />

Abstract<br />

Human activities often cause changes <strong>and</strong> homogenization in l<strong>and</strong>scape structure. To investigate<br />

the human impact on a tropical agricultural l<strong>and</strong>scape we mapped l<strong>and</strong> cover <strong>and</strong> small-scale<br />

linear l<strong>and</strong>scape elements in La Gamba (Costa Rica) <strong>and</strong> compared eight sections by different<br />

l<strong>and</strong>scape metrics. Rural sections clearly differed from forests, especially pasture-dominated<br />

sections including many linear l<strong>and</strong>scape elements <strong>and</strong> few big plantations. The largest <strong>and</strong><br />

most compact patches belonged to primary <strong>and</strong> secondary forests. Conversely, cultivated<br />

l<strong>and</strong>scapes were diverse comprising many small patches. Contrasting the results of other<br />

studies, most rural sections obtained higher fractal dimensions than forests, probably due to a<br />

higher density of linear l<strong>and</strong>scape elements. Natural l<strong>and</strong>scape elements such as live fences <strong>and</strong><br />

riparian vegetation which are supporting wildlife movement between forests are declining. Their<br />

protection is of major importance, particularly as the globally increasing cultivation of oil palms<br />

is significantly altering the countryside of La Gamba <strong>and</strong> many other tropical l<strong>and</strong> mosaics.<br />

Keywords: Costa Rica, Habitat fragmentation, La Gamba, L<strong>and</strong>scape metrics, L<strong>and</strong> use<br />

1. Introduction<br />

Nowadays biodiversity is highly threatened by human activities in all tropical regions of the<br />

world (Kappelle et al. 2003). The loss <strong>and</strong> fragmentation of tropical forests <strong>and</strong> rapid changes in<br />

l<strong>and</strong> use have great influence on the population dynamics of various native plant <strong>and</strong> animal<br />

species (e.g. Morera et al. 2005; Harvey et al. 2008b). By improving the ecological connectivity<br />

between forest patches <strong>and</strong> natural l<strong>and</strong>scape elements in agricultural areas the problem of<br />

habitat fragmentation can be alleviated (Morera et al. 2005). Therefore, the presence of natural<br />

l<strong>and</strong>scape elements, such as live fences, forest patches <strong>and</strong> gallery forests is of great importance<br />

for wildlife using cultivated areas (Harvey et al. 2005, 2008a; Seaman <strong>and</strong> Schulze 2009).<br />

The village “La Gamba” is situated at the edge of the Piedras Blancas National Park (southwest<br />

Costa Rica) <strong>and</strong> its agricultural area belongs to a zone that is important for wildlife exchange<br />

between forest areas. Since the founding of the village in the 1940s its economy has undergone<br />

several changes. Cattle breeding <strong>and</strong> rice production have always been important <strong>and</strong> large areas<br />

were deforested to gain farml<strong>and</strong>. From 1954 to 1961 bananas were the most common cash crop<br />

in La Gamba. During the great depression in the 1980s many plantations were ab<strong>and</strong>oned, rice<br />

fields <strong>and</strong> pastures remained as the most common forms of l<strong>and</strong> use (Klingler 2007). Nowadays<br />

* Corresponding author. Tel.: 0043-660-2568808<br />

Email address: t.hoebinger@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Höbinger et al. 2010. Impact of changing cultivation systems on the l<strong>and</strong>scape structure of La Gamba<br />

59<br />

rice production sharply decreased while the oil palm industry is on the rise. During the last<br />

decade many agricultural areas have been rapidly converted into oil palm plantations which are<br />

already the second most area consuming l<strong>and</strong> use type after pastures. The cultivation of oil<br />

palms is increasing globally, causing severe problems in many tropical agricultural systems <strong>and</strong><br />

leading to deforestation, loss of biodiversity, destruction of soils <strong>and</strong> alteration of traditional<br />

countryside (Fitzherbert et al. 2008). Until now in La Gamba new oil palm plantations are<br />

mainly replacing former rice fields or pastures <strong>and</strong> there has not been much deforestation.<br />

Nevertheless, this development profoundly affects the economic situation of La Gamba <strong>and</strong> has<br />

strong influence on the l<strong>and</strong>scape structure (Höbinger 2010).<br />

L<strong>and</strong>scape metrics are a useful too to analyze <strong>and</strong> monitor changes in l<strong>and</strong>scape pattern <strong>and</strong><br />

differences between l<strong>and</strong>scapes (Uuemaa et al. 2009). L<strong>and</strong>scape structure mirrors a wide range<br />

of ecological patterns <strong>and</strong> processes (Turner 2001). Several authors wrote about the use <strong>and</strong><br />

misuse of l<strong>and</strong>scape metrics (e.g. Li <strong>and</strong> Wu 2004) <strong>and</strong> tried to figure out appropriable sets of<br />

metrics for different scales <strong>and</strong> scientific questions (e.g. O’Neill et al. 1988; McGarigal <strong>and</strong><br />

Marks 1995; Botequilha Leitão et al. 2006; Cushman et al. 2008). A careful choice of metrics is<br />

essential to avoid redundancies <strong>and</strong> misleading results (McGarigal <strong>and</strong> Marks 1995; Schindler<br />

et al. 2008; 2009).<br />

This study analysis the l<strong>and</strong>scape pattern of the La Gamba area, <strong>and</strong> deviates ecological<br />

consequences of ongoing l<strong>and</strong> use change. It shall form a basis for further investigations <strong>and</strong><br />

action plans for biodiversity conservation in the area <strong>and</strong> other tropical l<strong>and</strong>scape mosaics.<br />

2. Methodology<br />

2.1 Study area <strong>and</strong> l<strong>and</strong> cover mapping<br />

The study area, the village La Gamba <strong>and</strong> surrounding forest areas, is situated in the southwest<br />

of Costa Rica (Golfo Dulce Region) <strong>and</strong> has an extend of 25.66 km² (8°41’ to 8°43’N <strong>and</strong> 83°9’<br />

to 83°13’W). To investigate the l<strong>and</strong>scape mosaic we mapped l<strong>and</strong> cover <strong>and</strong> small-scale linear<br />

l<strong>and</strong>scape elements (see Figure 1). The base for the mapping of the region was a “QuickBird 2”<br />

satellite image with a pixel size of 2.4 m for the multispectral channels (green, blue, red <strong>and</strong><br />

infrared) <strong>and</strong> 0.6 m for the panchromatic channel (La Gamba, Costa Rica, QuickBird scene<br />

052017330010_01_P001, 6/12/2007 © Digital Globe (2008), Distributed by Euroimage).<br />

To produce a vector map showing the l<strong>and</strong> cover, we used the program ArcView® (ESRI, Inc.,<br />

Redl<strong>and</strong>s, CA) <strong>and</strong> defined eleven l<strong>and</strong> cover categories: primary forest, secondary forest,<br />

shrubl<strong>and</strong>, riparian vegetation, fern-dominated vegetation, pastures, oil palm plantations, live<br />

fences <strong>and</strong> timber plantations, agriculture, settlements <strong>and</strong> roads <strong>and</strong> drainage ditches <strong>and</strong> rivers<br />

(including ponds). We used the statistics program R version 2.6.0 (R Development Core Team)<br />

to illustrate the results in the form of barplots.<br />

2.2 L<strong>and</strong>scape pattern analysis<br />

To analyze the l<strong>and</strong>scape pattern, the study area was divided into eight sections. All sections<br />

should represent relatively homogenous <strong>and</strong> characteristic zones of the area <strong>and</strong> were delineated<br />

in order to comprise mainly either forests or rural areas. Each section includes at least five of the<br />

eleven l<strong>and</strong> cover categories <strong>and</strong> is of compact shape (see Figure 1). Three of these sections<br />

were summarized as “forest sections” (Bolsa forest, Station forest <strong>and</strong> Bonito forest), being<br />

mainly covered by primary <strong>and</strong> secondary forest <strong>and</strong> five sections were summarized as “rural<br />

sections” (Bolsa, Bonito 1, Bonito 2, La Gamba <strong>and</strong> Station agriculture), being dominated by<br />

anthropogenic ecosystems (see Figure 1).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Höbinger et al. 2010. Impact of changing cultivation systems on the l<strong>and</strong>scape structure of La Gamba<br />

60<br />

To investigate the l<strong>and</strong>scape pattern of the area <strong>and</strong> to uncover differences among the eight<br />

l<strong>and</strong>scape sections, we used the software FRAGSTATS 3.3 (McGarigal <strong>and</strong> Marks 1995) to<br />

compute the l<strong>and</strong>scape metrics Patch Density (PD), Patch Area (AREA), Fractal Dimension<br />

(PAFRAC), Similarity Index (SIMI), Contagion Index (CONTAG) <strong>and</strong> Patch Richness Density<br />

(PRD). The different patch types were characterized by the class level metrics Patch Area<br />

(AREA), Fractal Dimension (PAFRAC), Euclidean Nearest Neighbor Distance (ENN) <strong>and</strong><br />

Edge Contrast (ECON). We applied the eight neighbor rule to guarantee that linear l<strong>and</strong>scape<br />

elements were identified as single patches (McGarigal <strong>and</strong> Marks 1995; Schindler et al. 2008).<br />

As the l<strong>and</strong>scape sections differed in size, we used only metrics st<strong>and</strong>ardized for area (e.g. Patch<br />

Density instead of the absolute number of patches).<br />

3 Results<br />

Primary vegetation covered 29% of the study area, secondary vegetation 35%, anthropogenic<br />

ecosystems 34% <strong>and</strong> water (rivers <strong>and</strong> ponds) 2% (see Figure 1). The most area consuming l<strong>and</strong><br />

use types of the agricultural l<strong>and</strong> mosaics were pastures (61 %) <strong>and</strong> oil palm plantations (31 %).<br />

Other l<strong>and</strong> use types (e.g. rice, cacao, bananas) covered only small areas (< 4%). All l<strong>and</strong>scape<br />

metrics clearly separated forests from rural sections. Generally, forest areas showed lower<br />

values of PD, PAFRAC <strong>and</strong> PRD <strong>and</strong> higher values of SIMI <strong>and</strong> CONTAG compared to rural<br />

areas (see Figure 2a). The differences between forests <strong>and</strong> rural areas were most evident for<br />

traditional pasture-dominated l<strong>and</strong>scapes, which typically consisted of many small patches of<br />

different type <strong>and</strong> included many linear l<strong>and</strong>scape elements such as live fences, streets or<br />

drainage ditches. Rural sections including few linear l<strong>and</strong>scape elements <strong>and</strong> big plantations or<br />

undivided pastures were characterized by metric values more similar to those of forest sections.<br />

Most rural sections of the study area had much higher fractal dimensions than forests. Patch<br />

types including lines showed the smallest patch areas <strong>and</strong> the highest fractal dimensions (see<br />

Figure 2b). Patches of primary forest had the biggest AREA <strong>and</strong> the lowest fractal dimensions.<br />

Secondary forests were smaller <strong>and</strong> had higher fractal dimensions. All other natural patch types<br />

such as riparian vegetation were of very small extent <strong>and</strong> more complex shaped. The<br />

comparison of AREA <strong>and</strong> PD showed that oil palm plantations consisted of bigger, but less<br />

numerous patches compared to pastures (see Figure 3). Primary forests had the biggest AREA,<br />

but the lowest PD. Secondary forests were characterized by a much smaller AREA <strong>and</strong> clearly<br />

higher PD. Riparian vegetation covered only smaller areas, but showed a relatively high PD.<br />

Categories including linear l<strong>and</strong>scape elements showed the smallest AREA <strong>and</strong> highest PD.<br />

4 Discussion<br />

For this study eight l<strong>and</strong>scape metrics were chosen with regard to the suggestions of other<br />

authors (Botequilha Leitão et al. 2006, Cushman et al. 2008, Schindler et al. 2008). The chosen<br />

metrics clearly distinguished forest <strong>and</strong> rural sections. <strong>Forest</strong> sections consisted of relatively<br />

few, big <strong>and</strong> compact shaped patches. Conversely, rural areas included more, smaller patches<br />

<strong>and</strong> were more diverse. Fractal dimensions were considerable high for rural areas. This can be<br />

caused by very high values of PAFRAC for the categories “settlement <strong>and</strong> road”, “drainage<br />

ditch <strong>and</strong> river” <strong>and</strong> “live fence <strong>and</strong> timber plantation” that included linear elements. This<br />

clearly demonstrates the importance of considering linear elements when assessing patch shape<br />

complexity.<br />

The inclusion of linear l<strong>and</strong>scape elements also has a strong influence on the values of<br />

l<strong>and</strong>scape metrics <strong>and</strong> dem<strong>and</strong>s a careful interpretation of the results. For example, the results of<br />

this study are not consistent with other studies that show that agriculture causes simple <strong>and</strong><br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Höbinger et al. 2010. Impact of changing cultivation systems on the l<strong>and</strong>scape structure of La Gamba<br />

61<br />

compact shaped l<strong>and</strong>scape patches (O’Neill et al. 1988). With exception of section Bonito 2 all<br />

rural sections had higher fractal dimensions than forests. Due to the inclusion of linear<br />

l<strong>and</strong>scape elements these high values indicate rather a high density of linear elements than a<br />

high complexity of the matrix. The comparison of AREA <strong>and</strong> PD clearly showed differences in<br />

the configuration of the patch types. Primary forests had the biggest extend, but comprised only<br />

few patches. Secondary forests were characterized by a much smaller AREA <strong>and</strong> clearly higher<br />

PD because most of these forests were formerly used as pastures. Riparian vegetation covered<br />

only smaller areas, but showed a relatively high PD because many small patches <strong>and</strong> strips of<br />

remnant riparian forests were present along the riversides.<br />

The expansion of oil palm plantations causes considerable changes in the l<strong>and</strong>scape matrix.<br />

Pasture-dominated parts were richer in ecologically valuable elements such as live fences <strong>and</strong><br />

riparian vegetation, while within <strong>and</strong> along oil palm plantations mainly ecologically futile<br />

elements such as roads <strong>and</strong> drainage ditches were found. Compared to pastures plantations had a<br />

bigger AREA, but lower PD which reflects that they are labor-intensive permanent cultures of<br />

great extend. The l<strong>and</strong> use map (see Figure 1) shows that, conversely to pastures, oil palm<br />

plantations were never divided <strong>and</strong> scarcely bordered by live fences. Hence, the expansion of<br />

these plantations involves the risk of a simplification of l<strong>and</strong>scapes <strong>and</strong> the loss of small natural<br />

l<strong>and</strong>scape elements in agricultural areas. This can lead to a significant reduction or shift of the<br />

biodiversity in tropical ecosystems affected by oil palm plantations.<br />

Oil palms are a globally rapidly exp<strong>and</strong>ing crop which has already replaced large areas of<br />

natural forest in many tropical countries such as Malaysia or Indonesia <strong>and</strong> this progress is still<br />

going on. These plantations entail many other problems such as habitat fragmentation, pollution,<br />

loss of biodiversity <strong>and</strong> soil destruction. Because oil palms are unsuitable habitat for most forest<br />

species they can act as severe barrier to animal movement (Fitzherbert et al. 2008). To maintain<br />

the exchange of plants <strong>and</strong> animals between the forest areas surrounding the village La Gamba,<br />

the conservation of live fences <strong>and</strong> other natural habitats is of great importance. As the mean<br />

patch area of pastures was relatively big (2.52 ha) <strong>and</strong> the density of live fences was rather low<br />

(20.0 m per ha farml<strong>and</strong>) more live fences could be established by dividing pastures into smaller<br />

paddocks (Höbinger 2010). This measure would improve the connectivity of forest patches,<br />

increase the tree cover within farml<strong>and</strong>, <strong>and</strong> provide a high conservation value of the<br />

agricultural l<strong>and</strong>scape mosaic in its unconnected gallery forests (Seaman <strong>and</strong> Schulze 2009).<br />

References<br />

Botequilha Leitão, A.B., Miller, J., Ahern, J. <strong>and</strong> McGarigal, K., 2006. Measuring <strong>L<strong>and</strong>scapes</strong>.<br />

A Planner's H<strong>and</strong>book. Isl<strong>and</strong> Press, Washington DC, 272 p.<br />

Cushman, S.A., McGarigal, K. <strong>and</strong> Neel, M.C., 2008. Parsimony in l<strong>and</strong>scape metrics: Strength,<br />

universality, <strong>and</strong> consistency. Ecological Indicators, 8: 691–703.<br />

Fitzherbert, E.B, Struebig, M.J., Morel, A., Danielsen, F., Brühl, C.A., Donald, P.F., <strong>and</strong><br />

Phalan, B., 2008. How will oil palm expansion affect biodiversity Trends in Ecology <strong>and</strong><br />

Evolution, 23: 538-545.<br />

Harvey, C.A., Guindon, C.F., Haber, W.A., Hamilton DeRosier, D. <strong>and</strong> Murray, K.G., 2008a.<br />

La importancia de los fragmentos de bosque, los árboles dispersos y las cortinas<br />

rompevientos para la biodiversidad local y regional: el caso de Monteverde, Costa Rica.<br />

In: Harvey C.A. <strong>and</strong> Saénz J.C. (Eds.). Evaluación y conservación de biodiversidad en<br />

paisajes fragmentados de Mesoamérica. Santo Domingo de Heredia, Cosa Rica: Instituto<br />

Nacional de Biodiversidad. Costa Rica. INBio: 289-325.<br />

Harvey, C.A., Sáenz, J.C. <strong>and</strong> Montero, J., 2008b. Conservación de la biodiversidad en<br />

agropaisajes de Mesoamérica: ¿que hemos aprendido y qué nos falta conocer In: Harvey<br />

C.A. <strong>and</strong> Saénz J.C. (Eds.). Evaluación y conservación de biodiversidad en paisajes<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Höbinger et al. 2010. Impact of changing cultivation systems on the l<strong>and</strong>scape structure of La Gamba<br />

62<br />

fragmentados de Mesoamérica. Santo Domingo de Heredia, Cosa Rica: Instituto<br />

Nacional de Biodiversidad. Costa Rica. INBio: 579-596.<br />

Harvey, C.A., Villanueva, C., Villacís, J., Chacón, M., Munoz, D., López, M., Ibrahim, M.,<br />

Gómez, R., Taylor, R., Martinez, J., Navas, A., Saenz, J., Sánchez, D., Medina, A.,<br />

Vilchez, S., Hernández, B., Perez, A., Ruiz, F., Lang, I., Sinclair, F. L., 2005.<br />

Contribution of live fences on the ecological integrity of agricultural l<strong>and</strong>scapes.<br />

Agriculture, Ecosystems <strong>and</strong> Environment, 111: 200–230.<br />

Höbinger, T., 2010. L<strong>and</strong> use, l<strong>and</strong>scape configuration <strong>and</strong> live fences in an agricultural area<br />

in southern Costa Rica: proposals for improving l<strong>and</strong>scape structure <strong>and</strong> establishment<br />

of biological corridors. Diploma Thesis. University of Vienna.<br />

Kappelle, M., Castro, M.A.H., Gonzáles, F. <strong>and</strong> Monge, H., 2003. Ecosystemas del Área de<br />

Conservación Osa (ACOSA) - Ecosystems of the Osa Conservation Area (ACOSA). Santo<br />

Domingo de Heredia, Cosa Rica: Instituto Nacional de Biodiversidad, INBio.<br />

Klingler, M., 2007. Wirtschafts- und siedlungsstruktureller W<strong>and</strong>el und seine Folgen in der<br />

Gemeinde La Gamba, Golfo Dulce Region, Costa Rica. Diploma Thesis. University of<br />

Innsbruck.<br />

Li, H. <strong>and</strong> Wu, J., 2004. Use <strong>and</strong> misuse of l<strong>and</strong>scape metrics. L<strong>and</strong>scape Ecology, 19: 389–<br />

399.<br />

McGarigal, K. <strong>and</strong> Marks, B., 1995. FRAGSTATS: Spatial Pattern Analysis Program for<br />

Quantifying L<strong>and</strong>scape Structure. USDA <strong>Forest</strong> Service General Technical Report PNW-<br />

GTR-351. Pacific Northwest Research Station, Portl<strong>and</strong>.<br />

Morera, C., Romero, M., Avenado, D. <strong>and</strong> Zuniga, A., 2005. Evaluación Socioambiental del<br />

área de amortiguamiento Este del Parque Piedras Blancas y La Reserva de Vida Silvestre<br />

Golfito. Revista Geográfica de América Central. Universidad Central.<br />

O'Neill, R.V., Krummel, J.R., Gardner, R.H., Sugihara, G., Jackson, B., DeAgelis, D.L., Milne,<br />

B.T., Turner, M.G., Zygmunt, B., Christensen, S.W., Dale, V.H. <strong>and</strong> Graham, R.L., 1988.<br />

Indices of l<strong>and</strong>cape pattern. L<strong>and</strong>scape Ecology, 1: 153–162.<br />

R Development Core Team, 2007. R: A Language <strong>and</strong> Environment for Statistical. Version<br />

2.6.0. Vienna, Austria: R Foundation for Statistical Computing. http://www.R-project.orgH<br />

(last accessed: 23. October 2009).<br />

Schindler, S., Poirazidis, K. <strong>and</strong> Wrbka, T., 2008. Towards a core set of l<strong>and</strong>scape metrics for<br />

biodiversity assessments: A case study from Dadia National Park, Greece. Ecology<br />

Indicators, 8: 502–514.<br />

Schindler, S., Kati, V., Von Wehrden, H., Wrbka, T. <strong>and</strong> Poirazidis, K., 2009. L<strong>and</strong>scape<br />

metrics as biodiversity indicators for plants, insects <strong>and</strong> vertebrates at multiple scales. In:<br />

Breu, J., Kozová, M. <strong>and</strong> Finka, M. (Eds.), European <strong>L<strong>and</strong>scapes</strong> in Transformation:<br />

Challenges for L<strong>and</strong>scape Ecology <strong>and</strong> Management. European IALE Conference 2009<br />

12-16 Jul 2009, Salzburg, Austria & Bratislava, Slovakia: 228-231.<br />

Seaman, B.S. <strong>and</strong> Schulze, C.H., 2009. The importance of gallery forests in the tropical<br />

lowl<strong>and</strong>s of Costa Rica for understorey forest birds. Biological Conservation, 143: 391-<br />

398.<br />

Turner, M.G, Gardner, R.H. <strong>and</strong> O’Neill, R.V., 2001. L<strong>and</strong>scape Ecology in Theory <strong>and</strong><br />

Practice. Pattern <strong>and</strong> Process. Springer-Verlag, New York.<br />

Uuemaa, E., Antrop, M., Roosaare, J., Marja, R. <strong>and</strong> M<strong>and</strong>er, Ü., 2009. L<strong>and</strong>scape Metrics <strong>and</strong><br />

Indices: An overview of Their Use in L<strong>and</strong>scape Research. Living Reviews in L<strong>and</strong>scape<br />

Research, 3, (2009), 1. http://www.livingreviews.org/lrlr-2009-1H (last accessed 11.<br />

November 2009).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Höbinger et al. 2010. Impact of changing cultivation systems on the l<strong>and</strong>scape structure of La Gamba<br />

63<br />

Figure 1: L<strong>and</strong> cover map <strong>and</strong> l<strong>and</strong>scape sections used for the l<strong>and</strong>scape pattern analysis.<br />

Figure 2: L<strong>and</strong>scape level metrics (a) <strong>and</strong> class level metrics (b). The values of SIMI, ENN <strong>and</strong> ECON<br />

represent the area-weighted means of the values of all single patches of the certain section (SIMI)/ of the<br />

certain l<strong>and</strong> cover type (ENN <strong>and</strong> ECON). For the explanation of the metric-acronyms, see section 2.2.<br />

Figure 3: Mean patch area of the l<strong>and</strong> cover types compared to their patch density.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Matos et al. 2010. Can lichen functional diversity be a good indicator of macroclimatic conditions<br />

64<br />

Can lichen functional diversity be a good indicator of macroclimatic<br />

conditions<br />

Paula Matos * , Pedro Pinho, Esteve Llop & Cristina Branquinho<br />

Universidade de Lisboa, Faculdade de Ciências, Centro de Biologia Ambiental, Campo<br />

Gr<strong>and</strong>e, Bloco C2, 5º Piso, Sala 37, 1749-016 Lisboa, Portugal<br />

Abstract<br />

Climate change is one of the greatest challenges facing conservation <strong>and</strong> it is predicted that its<br />

impact will be most significant in the Mediterranean region. In this work we proposed to study<br />

the effect of macroclimate on total lichen richness, abundance <strong>and</strong> on the proportion of lichen<br />

functional groups in order to find the best ecological indicators of macroclimatic conditions.<br />

The results showed that lichen functional-groups can be used as an indicator of the<br />

macroclimatic conditions at the l<strong>and</strong>scape level <strong>and</strong> showed to be better than total species<br />

richness or lichen abundance. By using three different functional groups we were able to<br />

observe shifts in the communities along the climatic gradient. Precipitation, evapotranspiration<br />

<strong>and</strong> relative air humidity were the variables that explained better the shifts from hygrophytic to<br />

xerophytic lichen communities. Lichen functional diversity showed to be a good c<strong>and</strong>idate for<br />

an ecological indicator of climate change.<br />

Keywords: lichen, functional-groups, species richness, abundance, climate change<br />

1. Introduction<br />

Climate change is predicted to affect many ecosystem services, including water shortages,<br />

increased risk of forest fires, northward shifts in the distribution of tree species, <strong>and</strong> losses of<br />

agricultural potential, which in Europe will be most significant in Mediterranean areas (Schroter<br />

et al. 2005). In particular in this region it is predicted an increase in drought conditions, due to<br />

higher temperature <strong>and</strong> reduced precipitation (IPCC 2007). There is a need to have tools for<br />

monitoring <strong>and</strong> even anticipating, the subtle changes that are or will occur due to climate change,<br />

both in space <strong>and</strong> time.<br />

Lichens are poikilohydric organisms that lack cuticle or roots, so they rely mainly on<br />

atmosphere for their water supply. Epiphytic lichens are considered one of the most sensitive<br />

groups to several changes that occur at ecosystem level, namely: air pollution (Branquinho et al.<br />

1999), ammonia (Pinho et al. 2009), l<strong>and</strong>-use (Pinho et al. 2008b) <strong>and</strong> microclimatic conditions<br />

(van Herk et al. 2002; Aptroot <strong>and</strong> van Herk 2007) due to their integrated sensitivity to changes<br />

in temperature <strong>and</strong>/or precipitation (Giordani <strong>and</strong> Incerti 2008).<br />

Although unconditionally considered good indicators of microclimatic alterations (van Herk<br />

2002; Aptroot <strong>and</strong> van Herk 2007), their use as indicators of macroclimatic conditions is not so<br />

unanimous. Moning et al. (2009), in the Bavarian <strong>Forest</strong> National Park, in southeastern<br />

Germany, found that total <strong>and</strong> threatened diversity of lichen species were mainly affected by<br />

forest structure, whereas macroclimatic factors were far less important, despite the steep average<br />

annual temperature gradient investigated, ranging from 4.2C to 7.8 °C. Other studies that<br />

focused on the impact of climate change on lichens suggested that lichens respond to global<br />

warming (Insarov et al. 1999). In the Netherl<strong>and</strong>s, arctic-alpine/boreo-montane species appear<br />

* Corresponding author.<br />

Email address: psmatos@fc.ul.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Matos et al. 2010. Can lichen functional diversity be a good indicator of macroclimatic conditions<br />

65<br />

to be declining, while (sub)tropical species are invading, independent of nutrient dem<strong>and</strong>s <strong>and</strong><br />

decreasing SO 2 emissions (van Herk et al. 2002). In Western Europe, the number of epiphytic<br />

species appears to be increasing rather than declining, as a result of global warming (Aptroot<br />

<strong>and</strong> van Herk 2007). Model predictions indicate major shifts in the distribution of lichen species<br />

(Ellis et al. 2007; Giordani <strong>and</strong> Incerti 2008).<br />

Despite many authors consider that climate change will affect lichens by shifting their<br />

communities, none of them tested this hypothesis explicitly, because these studies were<br />

performed considering either total species richness or individual species response. Recent<br />

studies (Giordani <strong>and</strong> Incerti 2008) focused on functional-diversity, but used posteriori selected<br />

guilds, that were based on the pattern of environmental variables. Because it was based on<br />

posteriori classification, it is strongly dependent on local communities, <strong>and</strong> thus cannot have a<br />

broader applicability. To overcome this problem <strong>and</strong> find a more universal indicator, we have<br />

made use of a priory classification of functional-diversity based on their humidity requirements.<br />

The use of lichen functional groups was shown to be better related with several environmental<br />

gradients than total biodiversity, at least for NH 3 air pollution (Pinho et al. 2009). Moreover,<br />

functional-groups of species have been successfully used to disentangle or analyze in<br />

simultaneous the influence of multiple environmental factors (Stofer et al. 2006; Pinho et al.<br />

2008a).<br />

In this work we propose to study the effect of macroclimate on total lichen richness, abundance<br />

<strong>and</strong> on the shifts of lichen functional groups in a regional area ranging from the more humid<br />

areas in coastal central Portugal to the SE of the Alentejo in semi-arid areas. This work intends<br />

to be a first approach to find the best ecological indicators of macroclimatic changes for<br />

Mediterranean areas.<br />

2. Methodology<br />

2.1 Study area<br />

The study was conducted in three different areas of continental Portugal, covering three<br />

different macroclimatic regions (Figure 1). The area A is a Quercus faginea wood located in the<br />

west central part of Portugal, within the Natural Park of Serra d’Aire e C<strong>and</strong>eeiros. This area<br />

belongs to the Mesomediterranean belt, from the Coastal Lusitano-Andalusian province (Rivas-<br />

Martinéz <strong>and</strong> Rivas-Saenz 2009). It has an annual average temperature that ranges from 15 to<br />

17.5 ºC <strong>and</strong> an annual average precipitation between 1400 <strong>and</strong> 1600 mm (averages from 1931 to<br />

1960, (IA 2010)). The area B is located in the southwest coast of Portugal, facing the Atlantic<br />

Ocean to the west. It’s a cork oak wood (Quercus suber) with annual average temperature<br />

between 16 <strong>and</strong> 17.5 ºC, <strong>and</strong> average annual precipitation between 600 <strong>and</strong> 1000 mm (averages<br />

from years 1931 to 1960 (IA 2010)). This area is in the Termomediterranean belt, within the<br />

Coastal Lusitano-Andalusian province (Rivas-Martinéz <strong>and</strong> Rivas-Saenz 2009). The area C is a<br />

holm oak wood (Quercus ilex) located southeast from the former area. This area belongs to the<br />

Mesomediterranean belt <strong>and</strong> is placed in the Mediterranean West Iberian province (Rivas-<br />

Martinéz <strong>and</strong> Rivas-Saenz 2009) <strong>and</strong> its climate is semi-arid warm, with annual average<br />

precipitation between 500 <strong>and</strong> 600 mm <strong>and</strong> average temperatures ranging between 16 <strong>and</strong> 17.5<br />

ºC.<br />

2.2 Biodiversity data collection <strong>and</strong> calculation of lichen-diversity variables<br />

Lichen diversity was sampled in 149 sites, 29 sampling sites in the area A, 77 sampling sites in<br />

the area B (Pinho et al. 2008a,b) <strong>and</strong> 43 sampling sites in the area C. The sampling was carried<br />

according to the method described in Asta et al. (2002).<br />

For each site we calculated several lichen-variables: i) total number of species; ii) LDV, lichen<br />

diversity value, which is the sum of all species frequency within the grid. Additionally LDV<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Matos et al. 2010. Can lichen functional diversity be a good indicator of macroclimatic conditions<br />

66<br />

was also calculated considering functional-groups, by dividing species according to humidity<br />

requirements considering species maximum classification in an available index (Nimis <strong>and</strong><br />

Martellos 2008). Therefore, species classified in the index with 1-2 were considered<br />

hygrophytes (LDVhygro), species classified with 3 were considered mesophytes (LDVmeso)<br />

<strong>and</strong> species classified with 4-5 xerophytes (LDVxero). Because we were dealing with different<br />

ecosystems the LDV values were used as relative values, i.e. the contribution (%) of each<br />

functional group to the total LDV.<br />

2.3 Climatic data <strong>and</strong> statistical analysis<br />

Lichen data was related with annual average values of macroclimatic series, available from<br />

Atlas do Ambiente (IA 2010). The variables selected were: precipitation (mm, 1931-1960);<br />

temperatre (ºC, 1931-1960); solar radiation (kcal/cm 2 , 1938-1970); real evapotranspiration (mm,<br />

1938-1970); insolation (h, 1931-1960); relative air humidity (%, 1931-1960). Values for each<br />

sampled site were estimated from the available maps. Correlations between lichen variables <strong>and</strong><br />

climatic variables were performed using Spearman rank order correlations considering the 149<br />

samples together.<br />

3. Results<br />

No correlation was found between the number of lichen species of each site <strong>and</strong> the long-term<br />

macroclimatic variables studied, except for relative air humidity (Table 2). However the total<br />

LDV value showed to be positively related not only with relative air humidity but also with<br />

insolation <strong>and</strong> solar radiation (Table 2).<br />

All the lichen functional diversity variables showed to be related with the long-term<br />

macroclimatic parameters studied, except the percentage of mesophytic LDV with temperature<br />

<strong>and</strong> relative air humidity (Table 2). The relative value of hygrophytic LDV was positively<br />

related with precipitation <strong>and</strong> real evapotranspiration, <strong>and</strong> negatively related with insolation<br />

(Figure 2). On the contrary, the relative value of xerophytic LDV showed to decrease with<br />

increasing precipitation <strong>and</strong> evapotranspiration, but was promoted by the increase in insolation<br />

(Figure 2).<br />

4. Discussion<br />

The results show that lichen functional-groups can be used as ecological indicators of the<br />

macroclimatic conditions in a regional area. We found that indicators based on lichen functional<br />

groups responded better to changes in macroclimatic conditions, than the total number of lichen<br />

species or its abundance (LDV). The fact that functional diversity responded more clearly to<br />

environmental gradients than total diversity was also found in other works but, for other<br />

environmental stresses, such as ammonia or atmospheric pollution (Pinho et al. 2008b; Pinho et<br />

al. 2009).<br />

The most important climatic factor driving the total richness <strong>and</strong> abundance (LDV) of lichens<br />

was the relative humidity. However, its abundance was also promoted by higher insolation.<br />

Although only some works (Heylen et al. 2005) showed this relation between species richness<br />

<strong>and</strong> relative air humidity, the poikylohydric nature of these organisms justifies the importance<br />

of this climatic factor. On the other h<strong>and</strong>, abundance was also related with insolation. Lichens<br />

need water to be physiologically active <strong>and</strong> they need light to photosynthesize. In this way, sites<br />

with higher relative air humidity <strong>and</strong> with a higher number of hours with sun will promote<br />

lichens activity <strong>and</strong> productivity, <strong>and</strong> ultimately lichens abundance. This hypothesis was also<br />

raised in a work in a tropical lowl<strong>and</strong> rain forest. Zotz <strong>and</strong> Winter (1994) suggested that the low<br />

abundance of macroclichens found there was mainly due to low light conditions, combined with<br />

high temperature.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Matos et al. 2010. Can lichen functional diversity be a good indicator of macroclimatic conditions<br />

67<br />

By using three different functional groups we were able to observe shifts in the communities<br />

along the climatic gradient. Precipitation, evapotranspiration <strong>and</strong> relative air humidity were the<br />

variables that explained better the shifts from hygrophytic to xerophytic lichen communities. In<br />

a Liguria case study, Girodani <strong>and</strong> Incerti (2008) found that 30% of its observed lichen flora<br />

was significantly correlated with yearly average temperature <strong>and</strong> rainfall patterns. This<br />

correlation allowed them to identify 3 guilds of species significantly sensitive to these climatic<br />

variables: species mainly related to cold-humid climate, species related to humid conditions but<br />

occurring in a wider range of temperatures <strong>and</strong> species occurring in meso to warm areas with<br />

humid to dry climate. Although not using the same functional groups, our results show a similar<br />

tendency, not evidenced by a group of individual species as in the former study, but with an a<br />

priori selected functional-group related to humidity requirements. Above 1400 mm in average<br />

of precipitation the hygrophytic community clearly dominated (>50%) over the mesophytic or<br />

the xerophytic ones, whereas below 600 mm it corresponded to only one third of the community.<br />

The xerophytic community increased clearly <strong>and</strong> consistently only when the precipitation was<br />

below approximately 600 mm. The mesophytic group showed an intermediate behavior. With<br />

increasing evapotranspiration the hygrophytic lichens showed a constant increase in relative<br />

average <strong>and</strong> also in variance. Again the xerophytic community responded more clearly only<br />

below 450 mm of evapotranspiration.<br />

The hygrophytic lichens responded negatively to increasing insolation, solar radiation <strong>and</strong><br />

temperature, although that was not as clear as observed for precipitation <strong>and</strong> evapotranspiration.<br />

This decrease in LDV of hygrophytic lichens was associated with an increase in xerophytic ones.<br />

These sets of variables (insolation, solar radiation <strong>and</strong> temperature) induce small changes in the<br />

median, <strong>and</strong> seem to change the variance of the response of the lichen communities.<br />

The group of functional indicators used/applied in this work can be very important in the current<br />

context of climate change. It was predicted for the Mediterranean region an increase in drought<br />

conditions, due to higher temperature <strong>and</strong> reduced precipitation (IPCC 2007). This work<br />

confirms that changes in macroclimatic conditions leads to changes in the functional structure of<br />

the lichen communities. Further studies are needed to evaluate the use of these indicators along<br />

time. Moreover, changes along smaller spatial scales should also be tested. The results of this<br />

work suggested that changes in sensitive lichen communities could be used as an indicator of<br />

macroclimatic changes in space.<br />

References<br />

Aptroot, A., van Herk, C.M., 2007. Further evidence of the effects of global warming on<br />

lichens, particularly those with Trentepohlia phycobionts. Env. Pollution, 146: 293-298.<br />

Asta, J., Erhardt, W., Ferretti, M., Fornasier, F., Kirschbaum, U., Nimis, P.L., Purvis, O.W.,<br />

Pirintsos, S., Scheidegger, C., van Haluwyn, C., Wirth, V., 2002. Mapping lichen<br />

diversity as an indicator of environmental quality. In: Nimis, P.L., Scheidegger, C.,<br />

Wolseley, P.A. (Eds.). Monitoring with Lichens - Monitoring Lichens. Nato Science<br />

Program, IV. The Netherl<strong>and</strong>s: Kluwer Academic Publisher: 273-279.<br />

Branquinho, C., Catarino, F., Brown, D.H., Pereira, M.J., Soares, A., 1999. Improving the use<br />

of lichens as biomonitors of atmospheric metal pollution. The Science of the Total<br />

Environment, 232: 67-77.<br />

Ellis, C.J., Coppins, B.J., Dawson, T.P., 2007. Predicted response of the lichen epiphyte<br />

Lecanora populicola to climate change scenarios in a clean-air region of Northern Britain.<br />

Biological Conservation, 135: 396-404.<br />

Giordani, P., Incerti, G., 2008. The influence of climate on the distribution of lichens: a case<br />

study in a borderline area (Liguria, NW Italy). Plant Ecology, 195: 257-272.<br />

van Herk, C.M., Aptroot, A., van Dobben, H.F., 2002. Long-tern monitoring in the Netherl<strong>and</strong>s<br />

suggests that lichens respond to global warming. Lichenologist, 34: 141-154.<br />

Heylen, O., Hermy, M., Schrevens, E., 2005. Determinants of cryptogamic epiphyte diversity in<br />

a river valley (Fl<strong>and</strong>ers). Biological Conservation, 126: 371-382.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Matos et al. 2010. Can lichen functional diversity be a good indicator of macroclimatic conditions<br />

68<br />

Insarov, G.E., Semenov, S.M., Insarova, I.D., 1999. A system to monitor climate change with<br />

epilithic lichens. Environmental Monitoring <strong>and</strong> Assessment, 55: 279–298.<br />

IPCC, I. P. o. C. C. 2007. Climate <strong>Change</strong> 2007: Synthesis Report: 73p.<br />

Lange, O.L., Kilian, E., Ziegler, H., 1986. Water-vapor uptake <strong>and</strong> photosynthesis of lichens -<br />

performance differences in species with green <strong>and</strong> blue-green-algae as phycobionts.<br />

Oecologia, 71: 104-110.<br />

Nimis, P.L. <strong>and</strong> Martellos, S., 2008: ITALIC - The Information System on Italian Lichens.<br />

Version 4.0. University of Trieste, Dept. of Biology, IN4.0/1<br />

(http://dbiodbs.univ.trieste.it/).<br />

Moning, C., Werth, S., Dziock, F., Bässler, C., Bradtka, J., Hothorn, T., Müller, J., 2009. Lichen<br />

diversity in temperate montane forests is influenced by forest structure more than climate.<br />

<strong>Forest</strong> Ecology <strong>and</strong> Management, 258: 745-751.<br />

Pinho, P., Augusto, S., Máguas, C., Pereira, M.J., Soares, A., Branquinho, C., 2008a. Impact of<br />

neighbourhood l<strong>and</strong>-cover in epiphytic lichen diversity: analysis of multiple factors<br />

working at different spatial scales. Environmental Pollution, 151: 414-422.<br />

Pinho, P., Augusto, S., Martins-Loução, M.A., Pereira, M.J., Soares, A., Máguas, C.,<br />

Branquinho, C., 2008b. Causes of change in nitrophytic <strong>and</strong> oligotrophic lichen species in<br />

a Mediterranean climate: Impact of l<strong>and</strong> cover <strong>and</strong> atmospheric pollutants. Environmental<br />

Pollution, 154: 380-389<br />

Pinho, P., Branquinho, C., Cruz, C., Tang, Y.S., Dias, T., Rosa, A.P., Maguas, C., Martins-<br />

Loucao, M.A., Sutton, M.A., 2009. Assessment of Critical Levels of Atmospheric<br />

Ammonia for Lichen Diversity in Cork-Oak Woodl<strong>and</strong>, Portugal, in: Sutton, M.A., Reis,<br />

S., Baker, S.M.H. (Eds.), Atmospheric Ammonia - Detecting Emission <strong>Change</strong>s <strong>and</strong><br />

Environmental Impacts. Dordrecht: Springer: 109-119.<br />

Rivas-Martínez, S. <strong>and</strong> Rivas-Sáenz, S., 1996-2009. Sistema de Clasificación Bioclimática<br />

Mundial. Centro de Investigaciones Fitosociológicas, España.<br />

(http://www.ucm.es/info/cif)<br />

Schroter, D., Cramer, W., Leemans, R., Prentice, I.C., Araujo, M.B., Arnell, N.W., Bondeau,<br />

A., Bugmann, H., Carter, T.R., Gracia, C.A., de la Vega-Leinert, A.C., Erhard, M., Ewert,<br />

F., Glendining, M., House, J.I., Kankaanpaa, S., Klein, R.J.T., Lavorel, S., Lindner, M.,<br />

Metzger, M.J., Meyer, J., Mitchell, T.D., Reginster, I., Rounsevell, M., Sabate, S., Sitch,<br />

S., Smith, B., Smith, J., Smith, P., Sykes, M.T., Thonicke, K., Thuiller, W., Tuck, G.,<br />

Zaehle, S., Zierl, B., 2005. Ecosystem service supply <strong>and</strong> vulnerability to global change<br />

in Europe. Science, 310: 1333-1337.<br />

Stofer, S., Bergamini, A., Aragón, G., Carvalho, P., Coppins, B.J., Davey, S., Dietrich, M., Edit<br />

Farkas, Kärkkäinen, K., Keller, C., Lökös, L., Lommi, S., Máguas, C., Mitchell, R.,<br />

Pinho, P., Rico, V.J., Truscott, A.-M., Wolseley, P.A., Watt, A., Scheidegger, C., 2006.<br />

Species richness of lichen functional groups in relation to l<strong>and</strong> use intensity. The<br />

Lichenologist, 38: 331–353.<br />

Zotz, G. <strong>and</strong> Winter, K., 1994. Photosynthesis <strong>and</strong> carbon gain of the lichen, leptogiumazureum,<br />

in a lowl<strong>and</strong> tropical forest. Flora, 189: 179-186.<br />

Table 2: Spearman rank order correlation coefficient values between total number of species, total LDV,<br />

relative LDV of lichen functional groups related to humidity requirements <strong>and</strong> the climatic variables.<br />

Marked (*) correlations are significant for P


P. Matos et al. 2010. Can lichen functional diversity be a good indicator of macroclimatic conditions<br />

69<br />

Figure 1: Location of the three areas studied.<br />

Figure 2: Relative functional diversity LDV related to water stress tolerance in response to the<br />

macroclimate variables – precipitation, real evapotranspiration <strong>and</strong> insolation.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.S. Meier et al. 2010. Projections of shifts in species distributions<br />

70<br />

Projections of shifts in species distributions: assessing the influence of<br />

macro-climate <strong>and</strong> local processes<br />

Eliane S. Meier * , Felix Kienast & Niklaus E. Zimmermann<br />

L<strong>and</strong> Use Dynamics, Swiss Federal Research Institute WSL, Zürcherstrasse 111,<br />

CH-8903 Birmensdorf, Switzerl<strong>and</strong><br />

Abstract<br />

It is currently unclear, whether tree species will be able to keep pace with the ongoing<br />

<strong>and</strong> likely accelerating shift in climate. Here, we estimated the influence of changing<br />

macro-climate <strong>and</strong> local processes on shifts in species distributions. First, we evaluated<br />

tree co-occurrence patterns in climate space <strong>and</strong> estimated the influence of these<br />

patterns on current <strong>and</strong> future species distributions using species distribution models.<br />

Second, we combined these models with migration rates from a process-model <strong>and</strong> a<br />

GIS path cost analysis, to estimate key processes influencing tree migration rates <strong>and</strong> to<br />

predict more realistic shifts in large-scale tree re-distributions. Our results showed, that<br />

biotic interactions mainly limited species distributions towards favorable growing<br />

conditions, while climate was directly limiting primarily where biotic interactions were<br />

low. L<strong>and</strong>scape fragmentation was further strongly limiting migration. In conclusion,<br />

this may lead to considerable time lags in range shifts <strong>and</strong> re-adjustment to new<br />

conditions during climate change, especially for late succession species.<br />

Keywords: biotic interactions, macro-climate, Europe, niche-based model, stress-gradient<br />

hypothesis.<br />

1. Introduction<br />

Macroclimate is hypothesized to play a key role in large-scale species distributions (Thuiller,<br />

Araújo <strong>and</strong> Lavorel 2004, Woodward 1987, Whittaker, Willis <strong>and</strong> Field 2001), <strong>and</strong> thus, the<br />

changing climate is expected to shift the distribution of suitable habitats for many species<br />

considerably (Root et al. 2003, Parmesan <strong>and</strong> Yohe 2003). This expectation is consistent with<br />

observations during Holocene climate changes, where species adapted their spatial distribution<br />

more or less rapidly to the changing climate conditions (Davis <strong>and</strong> Shaw 2001). For on-going<br />

<strong>and</strong> future climate change, however, it remains unclear whether species will be able to keep<br />

pace with accelerating rates of change (Iverson, Schwartz <strong>and</strong> Prasad 2004). On the one h<strong>and</strong>,<br />

new dispersal limitations emerge, such as anthropogenic l<strong>and</strong>scape-fragmentation that may<br />

cause the newly emerging suitable habitats to be insufficiently connected. On the other h<strong>and</strong>, the<br />

expected global warming is predicted to be one or more orders of magnitude faster than past<br />

climate change events (Solomon <strong>and</strong> Kirilenko 1997, Etterson <strong>and</strong> Shaw 2001) <strong>and</strong> the<br />

influence on range shifts by small-scale processes limiting species establishment, survival <strong>and</strong><br />

dispersal, such as biotic interactions (e.g. inter-specific competition, facilitation) or disturbances<br />

were so far often ignored in analysis on large-scale species responses to climate change (Caplat,<br />

An<strong>and</strong> <strong>and</strong> Bauch 2008, Brooker et al. 2007). According to the 'stress-gradient hypothesis',<br />

* Corresponding author. Tel.: +41 44 7392 560 - Fax: +41 44 7392 215<br />

Email address: eliane.meier@wsl.ch<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.S. Meier et al. 2010. Projections of shifts in species distributions<br />

71<br />

abiotic factors such as climate, topography <strong>and</strong> soil may primarily constrain species ranges<br />

under unfavourable growing conditions <strong>and</strong> here competition is generally low (Bertness <strong>and</strong><br />

Callaway 1994). Where abiotic conditions are more favourable, biotic interactions increase, <strong>and</strong><br />

competition may then, additionally to abiotic constraints, further determine species range limits<br />

(MacArthur 1972). The constraining effect of interlinked abiotic <strong>and</strong> biotic processes may not<br />

only be important when predicting current species distributions (Meier, 2010), but may be even<br />

more important when estimating species range shifts due to changing environmental conditions.<br />

Range shifts are mainly determined by the rate of plant establishment, growth <strong>and</strong> survival at a<br />

new location <strong>and</strong> dispersal abilities (Higgins et al. 2003), <strong>and</strong> are hence strongly linked to<br />

abiotic <strong>and</strong> biotic conditions. Moreover, because most modern l<strong>and</strong>scapes are highly fragmented,<br />

the density of individuals producing propagules is reduced <strong>and</strong> fewer <strong>and</strong> more distant sites for<br />

those propagules to colonize are available. This may further slow down migration rates (Iverson<br />

et al. 2004). Studying the interlinked effects of macroclimate, inter-specific competition <strong>and</strong><br />

l<strong>and</strong>scape-fragmentation may help to better estimate colonisable suitable habitats of species.<br />

2. Methodology<br />

2.1 Variation partitioning approach<br />

We examined the extent to which the variance in spatial patterns of species explained by species<br />

distribution models (SDMs) can be partitioned among abiotic <strong>and</strong> biotic predictors, <strong>and</strong> how<br />

these partitions depend on species characteristics. We fitted generalized linear models (GLMs)<br />

for 11 common tree species in Switzerl<strong>and</strong> using three different sets of predictor variables:<br />

biotic, abiotic, <strong>and</strong> the combination of both sets. We estimated by variance partitioning the<br />

proportion of the variance explained by biotic <strong>and</strong> abiotic predictors, jointly or independently.<br />

We then analyzed the linkage of these partitions with species traits using non-parametric tests<br />

(Mann–Whitney U-test <strong>and</strong> the Kruskal–Wallis test, depending on the number of classes<br />

differentiated).<br />

2.2 Co-occurrence patterns<br />

Further, we analysed correlations between the relative abundance of European beech<br />

(Fagus sylvatica) <strong>and</strong> three major competitor species (Picea abies, Pinus sylvestris <strong>and</strong><br />

Quercus robur) in environmental space, analyzing the variation in correlation along two major<br />

environmental gradients, namely summer rainfall <strong>and</strong> annual degree-day sum. In a next step, we<br />

projected these co-occurence patterns to geographic space. In a following spatial analysis, we<br />

used generalized additive models (GAM) to predict the spatial patterns of species abundances,<br />

<strong>and</strong> we evaluated from these models where <strong>and</strong> how much the simulated F. sylvatica<br />

distribution varied under current <strong>and</strong> future climates if potential competitor species were in- or<br />

excluded. For the analysis of co-occurrence we used ICP <strong>Forest</strong> level I data as well as climatic,<br />

topographic <strong>and</strong> edaphic variables as predictors for modelling the spatial distribution of species<br />

using SDMs.<br />

2.3 Implementing migration rates in SDMs<br />

In a final step, we calibrated SDMs using generalized linear models (GLMs) with ICP forest<br />

level I data <strong>and</strong> climatic, topographic, edaphic <strong>and</strong> l<strong>and</strong>-use variables to predict current <strong>and</strong><br />

future tree distributions, assuming either no or full migration when projecting to scenarios of<br />

future climate. Additionally, we combined our SDMs with an explicit simulation of dynamic<br />

tree migration rates (i.e., depending on interlinked effects of climate, inter-specific competition<br />

<strong>and</strong> l<strong>and</strong>scape connectivity). Dynamic migration rates were estimated from a process model<br />

(TreeMig; Lischke et al. 2004) using an intense sensitivity analysis of migration rates along<br />

gradients of climate, species competition <strong>and</strong> distance between suitable habitats in fragmented<br />

l<strong>and</strong>scapes. We combined the derived migration response with a GIS path cost analysis, to<br />

estimate climatically suitable <strong>and</strong> colonisable habitats <strong>and</strong> the rate of expected migration of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.S. Meier et al. 2010. Projections of shifts in species distributions<br />

72<br />

target species to reach such habitats given constraints of the l<strong>and</strong>scape (fragmentation, climate,<br />

competing species). By this, we were able to simulate realistic tree migration patterns across<br />

Europe at a comparably fine spatial resolution of 1km for the ongoing century.<br />

3. Results<br />

The variation partitioning approach showed, that over all climatic conditions, the joint<br />

contribution of biotic <strong>and</strong> abiotic predictors to explain deviance in SDMs was relatively small<br />

(~9%) compared to the contribution of each predictor set individually (~20% each). The<br />

influence of biotic predictors was higher for mid to late succession species than for early<br />

succession species.<br />

The results of the correlation analysis were in line with the ‘stress-gradient hypothesis’ for F.<br />

sylvatica: towards favourable growing conditions, its abundance was strongly linked with the<br />

abundance of its competitors, while this link weakened considerably towards unfavourable<br />

growing conditions. This resulted in a North-South <strong>and</strong> an elevation gradient throughout Europe,<br />

with stronger correlations in the South <strong>and</strong> at low elevations. The sensitivity analysis showed a<br />

potential spatial segregation of currently competing species with changing climate <strong>and</strong> a<br />

pronounced shift of zones where co-occurrence patterns may play a major role, but also a<br />

general reduction in interaction strength.<br />

Results from the sensitivity analysis of migration rates point to one effect that may help to<br />

explain aspects of the ‘stress gradient-hypothesis’; the higher the biotic interactions (i.e.,<br />

increased number of species occurring in a cell) towards species-specific favourable growing<br />

conditions, the more migration rates are limited. Climate seems to be directly limiting migration<br />

<strong>and</strong> finally constraining species’ distributions only where biotic interactions were low.<br />

L<strong>and</strong>scape fragmentation further lead to considerable time lags in range shifts for some species.<br />

In summary, the projected distributions by 2100 predicted from limiting migration rates<br />

dynamically with SDMs was rather in agreement with assumptions of “unlimited dispersal” for<br />

early succession species, while it matched rather with the assumption of “no migration” for<br />

medium <strong>and</strong> late succession species.<br />

4. Discussion<br />

The influence of biotic variables on SDM performance may indicate that community<br />

composition <strong>and</strong> other local abiotic factors or biotic processes strongly influence species<br />

distributions. However, the importance of species co-occurrence patterns for calibrating reliable<br />

species distribution models for use in climate effects projections does not seem to be equally<br />

crucial under all possible climatic conditions; in our correlation approach we were able to<br />

localise European areas (mostly low elevations, more southern part of the range) where<br />

inclusion of biotic predictors is recommended. Further we demonstrated, that implementing<br />

more realistic migration rates might substantially alter projections, <strong>and</strong> reduce the uncertainty in<br />

projections of species distributions under climate change scenarios.<br />

During recent climate change, many slow reproducing mid to late successional species<br />

may not be able to keep pace according to our analysis. Biotic interactions may mainly limit<br />

migration rates towards species-specific favourable growing conditions, while climate seems to<br />

be directly limiting primarily where biotic interactions are low. L<strong>and</strong>scape fragmentation may<br />

further lead to considerable time lags in range shifts, which may bring migration to a halt where<br />

migration is already relatively slow. Assessing the effects of interlinked processes such as<br />

climate, inter-specific interactions <strong>and</strong> l<strong>and</strong>scape-fragmentation on migration rates <strong>and</strong> species<br />

distributions in a dynamic <strong>and</strong> compound model reduces the uncertainty in projections of<br />

species distributions under climate change scenarios. If st<strong>and</strong>ard SDMs should be used in the<br />

future because of simplicity or where insufficient information on dynamic migration rates is<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.S. Meier et al. 2010. Projections of shifts in species distributions<br />

73<br />

available, one may consider for each species adequate migration assumptions (i.e. “no<br />

migration” for mid to late successional species <strong>and</strong> “unlimited migration” for early successional<br />

species), <strong>and</strong> may have to investigate more intensively the influence on range shifts by smallscale<br />

processes limiting species establishment, survival <strong>and</strong> dispersal, such as biotic interactions<br />

(e.g. inter-specific competition, facilitation) or disturbances, which were so far are often ignored<br />

in analysis on large-scale species distributions.<br />

Improved predictions of potential species distributions under future climates <strong>and</strong> in<br />

novel communities may assist strategies for sustainable forest management. Especially in this<br />

domain it is important to dynamically adapt management decisions to on-going climate changes,<br />

due to the long life span of trees. Tree life cycles take multiple decades to complete, <strong>and</strong> the rate<br />

at which trees can disperse <strong>and</strong> migrate, invade <strong>and</strong> form closed forests in areas that become<br />

climatically suitable is even slower. Incorporating such local processes by combining a dynamic<br />

model approach with large-scale SDMs will be crucial for a better underst<strong>and</strong>ing of how tree<br />

species interact in a changing climate. This is key to produce more accurate <strong>and</strong> detailed<br />

predictions of tree responses to climate change as are required by forest-management.<br />

References<br />

Bertness, M. D. & R. Callaway (1994) Positive interactions in communities. Trends in Ecology<br />

& Evolution, 9, 191-193.<br />

Brooker, R. W., J. M. J. Travis, E. J. Clark & C. Dytham (2007) Modelling species' range shifts<br />

in a changing climate: The impacts of biotic interactions, dispersal distance <strong>and</strong> the rate<br />

of climate change. Journal of Theoretical Biology, 245, 59-65.<br />

Caplat, P., M. An<strong>and</strong> & C. Bauch (2008) Interactions between climate change, competition,<br />

dispersal, <strong>and</strong> disturbances in a tree migration model. Theor Ecol, 209–220.<br />

Davis, M. B. & R. G. Shaw (2001) Range shifts <strong>and</strong> adaptive responses to Quaternary climate<br />

change. Science, 292, 673-679.<br />

Etterson, J. R. & R. G. Shaw (2001) Constraint to adaptive evolution in response to global<br />

warming. Science, 294, 151-154.<br />

Higgins, S. I., J. S. Clark, R. Nathan, T. Hovestadt, F. Schurr, J. M. V. Fragoso, M. R. Aguiar, E.<br />

Ribbens & S. Lavorel (2003) Forecasting plant migration rates: managing uncertainty for<br />

risk assessment. Journal of Ecology, 91, 341-347.<br />

Iverson, L. R., M. W. Schwartz & A. M. Prasad (2004) How fast <strong>and</strong> far might tree species<br />

migrate in the eastern United States due to climate change <strong>Global</strong> Ecology <strong>and</strong><br />

Biogeography, 13, 209-219.<br />

Lischke, H., N. E. Zimmermann, J. Bolliger, S. Rickebusch & T. J. Loffler. 2004. TreeMig: A<br />

forest-l<strong>and</strong>scape model for simulating spatio-temporal patterns from st<strong>and</strong> to l<strong>and</strong>scape<br />

scale. In International Conference of the International-Environmental-Modelling-<strong>and</strong>-<br />

Software-Society, 409-420. Osnabruck, GERMANY: Elsevier Science Bv.<br />

MacArthur, R. H. 1972. Geographical ecology: patterns in the distribution of species.<br />

Parmesan, C. & G. Yohe (2003) A globally coherent fingerprint of climate change impacts<br />

across natural systems. Nature, 421, 37-42.<br />

Root, T. L., J. T. Price, K. R. Hall, S. H. Schneider, C. Rosenzweig & J. A. Pounds (2003)<br />

Fingerprints of global warming on wild animals <strong>and</strong> plants. Nature, 421, 57-60.<br />

Solomon, A. M. & A. P. Kirilenko (1997) Climate change <strong>and</strong> terrestrial biomass: what if trees<br />

do not migrate <strong>Global</strong> Ecology <strong>and</strong> Biogeography Letters, 6, 139-148.<br />

Thuiller, W., M. B. Araújo & S. Lavorel (2004) Do we need l<strong>and</strong>-cover data to model species<br />

distributions in Europe Journal of Biogeography, 31, 353-361.<br />

Whittaker, R. J., K. J. Willis & R. Field (2001) Scale <strong>and</strong> species richness: towards a general,<br />

hierarchical theory of species diversity. Journal of Biogeography, 28, 453-470.<br />

Woodward, F. I. 1987. Climate <strong>and</strong> plant distribution. Cambridge University Press. London.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Moutahir et al. 2010. Application of remote sensing to assess wildfire impact<br />

74<br />

Application of remote sensing to assess the wildfire impact on the<br />

natural vegetation recovery <strong>and</strong> l<strong>and</strong>scape structure in the<br />

Mediterranean forest of South eastern Spain<br />

H. Moutahir 1* , J.F. Bellot 1 , A.Bonet 1 , M.J. Baeza 2 & I. Touhami 1<br />

1 Departamento de Ecología, Universidad de Alicante, Ap. 99, 03080 Alicante, Spain<br />

2 Fundación CEAM, Parque Tecnológico Paterna, C/Charles Darwin 14, 46980<br />

Valencia, Spain<br />

Abstract<br />

<strong>Forest</strong> fires in the Mediterranean zone cause major disturbances in l<strong>and</strong>scape structure<br />

especially in the wooded stratum, since the recovery of the initial state of the natural vegetation<br />

is a complex <strong>and</strong> very slow process. This process, depending on many factors, passes through<br />

several stages leading to changes in the l<strong>and</strong>scape structure through time. In this research,<br />

several L<strong>and</strong>sat TM <strong>and</strong> ETM+ images were used to assess the impact of wild fire in the<br />

Mediterranean forests of south eastern Spain over the past 25 years. The use of the Normalized<br />

Difference Vegetation Index (NDVI) for mapping the burnt zones <strong>and</strong> indicators of l<strong>and</strong>scape<br />

heterogeneity allowed us to analyze the degree of l<strong>and</strong>scape change in these forests caused by<br />

the fires. In the majority of the observed zones the natural vegetation did not completely recover<br />

over this time period. The alteration in l<strong>and</strong>scape structure <strong>and</strong> its floristic composition depends<br />

on the type of vegetation <strong>and</strong> its maturity before the fire, topography, microclimates, the climate<br />

general, <strong>and</strong> finally, l<strong>and</strong> use <strong>and</strong> management.<br />

Keywords: <strong>Forest</strong> fire, Vegetation recovery, L<strong>and</strong>scape structure, Remote sensing, NDVI<br />

1. Introduction<br />

Situated on the Mediterranean coast in the South East of Spain, The Valencian Community is<br />

one of the most affected areas by the wildfires. In 1994 a total of 138404 ha of wooded <strong>and</strong> non<br />

wooded area were burnt which represents a third of the burnt area in Spain. Due to their<br />

magnitude wildfires are considered one of the most important driving forces determining the<br />

current forest l<strong>and</strong>scape in the Mediterranean basin (Naveh 1994) <strong>and</strong> several authors speak<br />

about a homogenization of the l<strong>and</strong>scape after fires. These fires affect l<strong>and</strong>scape patterns which<br />

can affect the ecological processes (Turner, 1989).<br />

The satellite images thanks to their spatial <strong>and</strong> temporal resolution allow the assessment of<br />

wildfires impact on the l<strong>and</strong>scape. In this way we carry out this research using a several L<strong>and</strong>sat<br />

TM <strong>and</strong> ETM+ images of a burnt area in the South East of the Valencian Community.<br />

* Corresponding author. Tel.: +34 644474238 - Fax: +34 965909832<br />

Email address: hassane_moutahir@yahoo.fr<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Moutahir et al. 2010. Application of remote sensing to assess wildfire impact<br />

75<br />

2. Methodology<br />

2.1 Study area<br />

The chosen area is located about 70 km south east of the city of Valencia (figure 1). The<br />

dominant vegetation type covering the l<strong>and</strong>scape are the evergreen shrubl<strong>and</strong>s (with abundance<br />

of Quercus coccifera) <strong>and</strong> pine woodl<strong>and</strong>s (Pinus halepensis) with different degrees of<br />

development <strong>and</strong> species compositions (Abdel Malak, 2009). This area suffered several fires in<br />

the last 25 years.<br />

2.2 Methodology<br />

In this research we follow the same methodology used by Chuvieco (1996, 2007) in his study to<br />

measure the l<strong>and</strong>scape structure using remote sensing images. Chuvieco based his research on<br />

only two images in order to compare the l<strong>and</strong>scape structure before <strong>and</strong> after the fire. However,<br />

for this research twelve L<strong>and</strong>sat TM <strong>and</strong> ETM+ images were used in order to obtain a more<br />

accurate assessment of the evolution of l<strong>and</strong>scape structure (table 1).<br />

The twelve satellite images used have been downloaded for free from the USGS <strong>Global</strong><br />

Visualization Viewer (http://glovis.usgs.gov/.). We did not perform any geometric correction to<br />

these images because the level of their correction was already satisfactory. Nevertheless, in<br />

order to compare them a radiometric normalization was carried out using the Pseudo-Invariant<br />

Feature Normalization method (Scott et al., 1988).<br />

To detect the burnt area we used a combination of the NDVI b<strong>and</strong>s calculated from the 12<br />

images <strong>and</strong> stretched between 0 <strong>and</strong> 200. Displaying the NDVI b<strong>and</strong> of the first year (1984) in<br />

both the red <strong>and</strong> the blue channels <strong>and</strong> displaying the rest of the years in the green channel, we<br />

can detect the burnt zone as a magenta colored area due to the low values of NDVI in the green<br />

channel.<br />

In order to quantify the l<strong>and</strong>scape structure two kinds of measures were applied:<br />

-measures applied to the NDVI images as continuous values: (the st<strong>and</strong>ard deviation through a<br />

profile) which measures the spatial contrast of the image.<br />

-measures applied to intervals of NDVI which measure the spatial structure of a territory.<br />

3. Results<br />

Analyzing only two images, one image before <strong>and</strong> one after the fire, we obtained the same<br />

results as Chuvieco did in 1996. In both profiles (10.5km <strong>and</strong> 4.5km of length) chosen within<br />

two different burnt areas at different times (the first area was burnt in 1986 <strong>and</strong> the second one<br />

was burnt in 1991), we observed that the st<strong>and</strong>ard deviation decrease after the fire: 9% in the<br />

first profile <strong>and</strong> 3% in the second profile (table 2&3; figure 2&3). This result indicates that the<br />

fire tends to homogenize the l<strong>and</strong>scape. However, analyzing the rest of images we observed that<br />

the NDVI <strong>and</strong> the st<strong>and</strong>ard deviation increase with time while the vegetation recover (table<br />

2&3).<br />

The second type of l<strong>and</strong>scape structure measures was applied to a window of 1473.7 ha of an<br />

area burnt in 1986. These measures were applied to classified images of 10 intervals of NDVI<br />

obtained from an automatic segmentation. The mean area of patches <strong>and</strong> the index of patch<br />

dominance increased while the number of patches, the density of patches <strong>and</strong> the mean diversity<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Moutahir et al. 2010. Application of remote sensing to assess wildfire impact<br />

76<br />

decreased after the fire (table 4). Nevertheless, it does not have the same tendency over all the<br />

observed time period.<br />

4. Discussion<br />

All results obtained in the year after the fire lead to one conclusion: the wildfires tend to<br />

homogenize the l<strong>and</strong>scape. The same conclusion was obtained for chuvieco in his study in 1996<br />

but we cannot conclude that the homogenization is forever because all indexes measured in this<br />

study change with time <strong>and</strong> there are differences in space.<br />

The difference in the NDVI values over the time period observed between the two profiles<br />

studied is due to the fact that the first profile was chosen inside an area burnt firstly in 1979 <strong>and</strong><br />

secondly in 1986. The second one was chosen inside another area burnt only one time in 1991.<br />

In the first case the initial vegetation was composed for mattorrals <strong>and</strong> shrubl<strong>and</strong>s because of<br />

the first fire <strong>and</strong> in the second case the initial vegetation was a wooded stratum that is what<br />

explains the differences in the vegetation recovery after the two fires. In the first case the pine<br />

forest can disappear (Baeza <strong>and</strong> al, 2007). The initial vegetation <strong>and</strong> the variation on the<br />

precipitations (table 1) can together explain the variability of the NDVI through time.<br />

References<br />

Abdel Malak, D., 2009. Fire regimes <strong>and</strong> post-fire regeneration in the Eastern Iberian<br />

Peninsula using GIS <strong>and</strong> remote sensing techniques. PhD thesis. University of<br />

Valencia.<br />

Baeza, M.J., Valdecantos, A., Alloza, J.A. & Vallejo, V.R, 2007. Human disturbance <strong>and</strong><br />

environmental factors as drivers of long-term post-fire regeneration patterns in<br />

Mediterranean forests. Journal of Vegetation Science 18: 243-252.<br />

Chuvieco, E., 1996. Empleo de imágenes de satélite para medir la estructura del paisaje:<br />

análisis cuantitativo y representación cartográfica. Serie Geográfica, vol. 6, pp. 131-<br />

147<br />

Chuvieco, E., 2007. Teledetección Ambiental. La observación de la tierra desde el espacio. 3º<br />

Edición. Ariel. Barcelona, 586 p.<br />

Naveh, Z., 1994. The role of fire <strong>and</strong> its management in the conservation of the Mediterranean<br />

ecosystems <strong>and</strong> l<strong>and</strong>scapes. In: Moreno, J.M. & Oechel, W. (eds.) The role of fire in<br />

Mediterranean type ecosystems. Springer-Verlag, New York, NY, US: 163-186.<br />

Schott, J.R., C. Salvaggio, <strong>and</strong> W.J. Volchok, 1988. Radiometric scene normalization using<br />

pseudoinvariant features, Remote Sensing of Environment. 26: l-16.<br />

Turner, M.G., 1989. L<strong>and</strong>scape ecology: the effect of patterns on process. Ann. Rev. of<br />

Ecological Systems 20: 171-197<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Moutahir et al. 2010. Application of remote sensing to assess wildfire impact<br />

77<br />

SPAIN<br />

Velencia<br />

Figure1: Geographic situation of the studied area<br />

Table1: Images dates <strong>and</strong> cumulative precipitation<br />

Cumulative Precipitation<br />

(mm/10 Months (Sept-June))<br />

Image Date<br />

Bañares station<br />

28/07/1984 302.9<br />

23/06/1986 394.9<br />

26/06/1987 880.7<br />

06/09/1990 820.2<br />

20/04/1992 364<br />

29/06/1994 373<br />

21/07/1999 257.4<br />

19/06/2002 497.4<br />

08/07/2003 379.6<br />

18/05/2005 418<br />

03/07/2007 453.5<br />

24/07/2009 730.8<br />

Table 2: NDVI mean values <strong>and</strong> st<strong>and</strong>ard deviation through a profile for the fire of 1986<br />

1984 1986* 1987 1990 1992** 1994 1999 2002 2003 2005 2007 2009<br />

Mean 136.9 126.9 132.5 133.7 134.3 132.3 141.7 139.8 134.2 137.2 134.8 134.6<br />

St<strong>and</strong>ard<br />

deviation 3.3 3.0 3.1 4.1 3.3 2.8 4.6 4.9 3.5 3.7 3.8 3.9<br />

*year of the fire, **Image captured in April<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Moutahir et al. 2010. Application of remote sensing to assess wildfire impact<br />

78<br />

Figure 2: NDVI value through a profile on Matorral cover before <strong>and</strong> after the fire of 1986<br />

Table 3: NDVI mean values <strong>and</strong> st<strong>and</strong>ard deviation through a profile for the fire of 1991<br />

1984 1986 1987 1990 1992* 1994 1999 2002 2003 2005 2007 2009<br />

Mean 143.4 140.5 142.1 141.0 117.9 132.1 144.4 142.9 135.0 138.5 136.9 137.4<br />

St<strong>and</strong>ard<br />

deviation 4.3 4.0 4.2 3.8 3.7 3.2 5.7 4.5 3.4 3.9 4.3 4.4<br />

*year of the fire is 1991<br />

Figure 3: NDVI value through a profile on wooded stratum before <strong>and</strong> after the fire of 1991<br />

Table 4: measures applied to 10 intervals of NDVI<br />

1984 1986 1987 1990 1992 1994 1999 2002 2003 2005 2007 2009<br />

Number of patch 7082 5707 6146 7228 7152 6747 7299 7328 6906 6962 6844 7306<br />

Mean area (pixels of<br />

30m) 2.6 3.2 3.0 2.5 2.5 2.7 2.5 2.5 2.6 2.6 2.7 2.5<br />

Density of patch /ha 130.1 104.8 112.9 132.8 131.4 124.0 134.1 134.6 126.9 127.9 125.7 134.2<br />

Mean diversity 2.2 2.2 1.9 1.8 1.6 1.5 1.4 1.4 1.2 1.2 1.1 1.0<br />

Dominance 0.1 0.2 0.4 0.5 0.7 0.9 1.0 0.9 1.1 1.1 1.2 1.3<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Palaghianu 2010. The use of Voronoi tessellation to characterize sapling populations<br />

79<br />

The use of Voronoi tessellation to characterize sapling populations<br />

Ciprian Palaghianu *<br />

<strong>Forest</strong>ry Faculty, “Stefan cel Mare” University of Suceava, Romania<br />

Abstract<br />

The area potentially available to an individual plant represents a concept extensively used in<br />

population ecology but it has fewer implementations in forest research. In this paper I use a<br />

Voronoi tessellation in order to determine area potentially available to a sapling. The Voronoi<br />

polygons were used to characterize spatial pattern of sapling distribution as well as the<br />

competition relations between the individuals. Mathematically, the Voronoi tessellation<br />

represents one of the best solutions to determine neighbouring competitors of a tree. The area of<br />

Voronoi cells is frequently connected to biometrical attributes <strong>and</strong> the growth of the saplings.<br />

Furthermore, analyzing the Voronoi tessellation of a sapling population can indicate the spatial<br />

pattern of the saplings. It is considered that weighted Voronoi polygons may be more fitted for<br />

assessing sapling relationships but it is more difficult to implement such specific algorithms.<br />

Keywords: Voronoi tessellation, area potentially available, spatial pattern, sapling populations<br />

1. Introduction<br />

Researches used many times mathematical <strong>and</strong> especially geometrical techniques in their effort<br />

to explain individual competition.<br />

The area potentially available (APA) concept represents an uncommon, but rather promising<br />

approach, introduced in plant ecology by Brown (1965). The same concept was independently<br />

developed by Mead (1966), but early investigations in the field of plants growing space were<br />

conducted also by Konig, mentioned in his book „Die Forst-Mathematik” (1835).<br />

From the biological point of view, APA generally defines the area used by an individual to<br />

access vital resources, the available area for a plant to satisfy its needs in water, nutrients <strong>and</strong><br />

light. So APA is very appealing to researchers interested in growth modelling, in their effort to<br />

solve an everlasting problem: “Do trees grow faster because they are larger Or they are larger<br />

because they have been growing faster” (Wichmann, 2002; Garcia, 2008).<br />

Considering the difficulty of the analysis there are few researches using this approach (Mark,<br />

Esler, 1970; Moore et al., 1973; Mercier, Baujard, 1997). Smith (1987) considers that this<br />

approach is ignored or even avoided due to misapprehend of APA geometrical foundation <strong>and</strong><br />

computing difficulties. The late period is well-known for its computer development <strong>and</strong> also<br />

numerous <strong>and</strong> various algorithms were produced. So, the APA re-enters in researcher’s<br />

attention as a promising investigation tool.<br />

The APA was used to solve not only competition issues but also mortality <strong>and</strong> dynamics of<br />

seedlings (Owens, Norton, 1989) or spatial pattern (Mercier, Baujard, 1997). Regarding spatial<br />

pattern, Garcia (2008) considers the interaction between neighbouring growing areas as a result<br />

of autocorrelation. Two neighbours which are closer than average, will both have APA<br />

undersized values <strong>and</strong> vice versa. Winsauer <strong>and</strong> Mattson (1992) have mentioned some<br />

advantages to make use of APA in forest researches – potentially available areas are not<br />

intersecting each other, there are sensitive to population dynamics <strong>and</strong> they are correlated with<br />

* Ciprian Palaghianu. Tel.:+40745614487<br />

Email address: cpalaghianu@usv.ro<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Palaghianu 2010. The use of Voronoi tessellation to characterize sapling populations<br />

80<br />

growth rates. This final remark represents the key aspect of APA utilisation as a competition<br />

evaluation tool because if an individual has a large APA, the competition pressure will affect it<br />

less. There is, of course, a drawback – the APA is based exclusively on the position of the<br />

individual <strong>and</strong> not on its biometrical attributes. That’s why it is called “potentially”.<br />

The objective of this study is to elucidate what kind of information APA can offer regarding<br />

sapling populations. Can APA characterize the relationships between saplings A subsidiary<br />

objective is producing software tools for Voronoi analysis.<br />

2. Methodology<br />

The area potentially available of a tree has experienced different forms of interpretation <strong>and</strong> use,<br />

analogous to Brown concept. For example, Staebler (1951), Bella (1971) <strong>and</strong> Moore (et al.,<br />

1973) used in their researches a similar concept named “influence zone”. Polygon areas were<br />

used as descriptive tool of spatial plant arrangement or as predictive tool of plant performance.<br />

The most correct interpretation remains although the one based on the mathematical concept of<br />

space partitioning using Voronoi tessellation. So it is generally admitted that APA of an<br />

individual is equivalent to area of the Voronoi polygon which is associated to that individual.<br />

In the bi-dimensional space, a Voronoi polygon of an element includes all the points closer to<br />

that specific element than to any other element. The edges of a polygon contain the points<br />

located at equal distance from two elements. The vertices of such a polygon are equally located<br />

from minimum three generating elements. Considering these properties, two elements are<br />

considered to be neighbours if their associated Voronoi polygons share an edge.<br />

It is quite difficult to obtain a Voronoi partitioning for a large set of points, that’s why this<br />

process is frequently done using specific algorithms <strong>and</strong> a computer. Algorithms were poorly<br />

optimized <strong>and</strong> the computers were very slow few decades ago, so the process was not pleasant<br />

<strong>and</strong> quick. The last years arrived with great improvements regarding the algorithms <strong>and</strong> the<br />

computer instruments, giving a new chance to Voronoi based applications.<br />

2.1 Developing the software tools<br />

In order to study the area potentially available to saplings I have developed specific software<br />

tools, using Microsoft Visual Basic. For the first tool, called VORONOI, I have used an<br />

algorithm presented by Ohyama (2008) with O(n2) complexity. VORONOI is st<strong>and</strong>-alone<br />

software which is drawing the Voronoi diagrams using as input data the saplings coordinate<br />

placed in a spreadsheet. The user can obtain information regarding sapling neighbours by<br />

diagrams analyses. The Voronoi tessellation represents a natural method to select neighbouring<br />

trees, a difficult issue in assessing competition indices. The diagrams also offer information<br />

about spatial pattern of saplings – it’s easier to determine if a pattern is aggregate or uniform.<br />

The second software tool, named ARIA VORONOI, computes the area of each Voronoi<br />

polygon. These areas, equivalent to APA values, might be used as competition or aggregation<br />

index. Small values of APA might indicate competition pressure <strong>and</strong> great values of APA<br />

coefficient of variation might indicate aggregation of saplings for the analyzed plot.<br />

This software was also developed in Microsoft Visual Basic. The input data represents the<br />

saplings Cartesian coordinates, extracted from a spreadsheet. The programme computes area of<br />

Voronoi polygons <strong>and</strong> several statistic indicators - the average, st<strong>and</strong>ard deviation <strong>and</strong><br />

coefficient of variation of APA values. It is generated a grid <strong>and</strong> each cell of the grid is analyzed<br />

to asses which the generator point (sapling) is. The user can choose a grid size step in order to<br />

increase accuracy of determining APA values.<br />

It was taken into account the edge effect, so the saplings with incomplete APA were eliminated<br />

from the analyses. User can specify a value for the buffer zone – in this way the APA is<br />

calculated only for the saplings located in the core area, even if the APA extends outside the<br />

core area. If the buffer zone is too small, in some exceptional cases, there might be saplings<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Palaghianu 2010. The use of Voronoi tessellation to characterize sapling populations<br />

81<br />

located in the core area with incomplete APA (Figure 1). The algorithm computes also the<br />

convex hull <strong>and</strong> all the points (saplings) located on the convex hull are eliminated. The<br />

recommended size of the buffer zone is the average distance between neighbours corrected with<br />

the coefficient of variation (20 cm in this study).<br />

2.2 Material <strong>and</strong> analyses<br />

Figure 1: Voronoi tessellation - the buffer zone <strong>and</strong> the convex hull of a plot<br />

The study area is located in Flămânzi <strong>Forest</strong> District, parcel 50A, near Cotu, a small settlement<br />

situated in Botoşani County, Romania. The topography is almost flat, with a slope average of 2-<br />

3% <strong>and</strong> the altitude is around 140 meters. The area of the st<strong>and</strong> studied is 21.5 hectares <strong>and</strong> the<br />

species composition consists of 30% sessile oak, 20% oak, 30% common hornbeam, 10%<br />

small-leaved linden <strong>and</strong> 10% common ash. The area is regenerated naturally <strong>and</strong> the<br />

regeneration gaps were created in 2001-2002 <strong>and</strong> were enlarged in 2007. Within this st<strong>and</strong> a 2.5<br />

hectare homogenous area covered in saplings was selected for further investigation.<br />

I installed a network of ten permanent rectangular sampling plots (7 x 7 m) where I measured<br />

the characteristics of all saplings <strong>and</strong> seedlings. I used a GPS receiver in order to record the<br />

coordinates of the centre of each plot <strong>and</strong> I labelled every sapling <strong>and</strong> seedling inside the plot.<br />

The features of 7253 individuals were determined. The attributes assessed are: species, location<br />

of the individuals (x, y Cartesian coordinates), diameter, total height, crown insertion height,<br />

two crown diameters along the directions of axes <strong>and</strong> the latest annual height growth.<br />

The analyses were based on areas of generated Voronoi polygons. There were studied the<br />

correlations between APA values <strong>and</strong> the main structural attributes (dimensional attributes,<br />

density), height growth <strong>and</strong> competition indices – Hegyi (1974) <strong>and</strong> Schutz (1989).<br />

APA it was also used as a potentially spatial pattern indicator. There are several studies<br />

(Mercier, Baujard, 1997; Garcia, 2008) that point out there is a relation between APA values or<br />

APA coefficient of variation <strong>and</strong> the spatial pattern of a population of mature trees. This fact<br />

might be relevant to sapling populations, too.<br />

In this case it was analyzed the APA coefficient of variation for each plot in relation with<br />

Morisita (1962) <strong>and</strong> Clark-Evans (1954) spatial pattern indicators. APA coefficient of variation<br />

might be an indicator of spatial distribution. In order to find a relationship between aggregation<br />

<strong>and</strong> APA coefficient of variation values I have used a statistical test to establish if there is a<br />

significant deviation from Poisson spatial distribution (from the complete spatial r<strong>and</strong>omness -<br />

CSR hypothesis). I have generated 19 Monte-Carlo simulations for each plot, using SpPack<br />

software (Perry, 2004) to simulate a CSR distribution for the same area <strong>and</strong> the same number of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Palaghianu 2010. The use of Voronoi tessellation to characterize sapling populations<br />

82<br />

saplings. The extreme values of APA coefficient of variation produced the 95% confidence<br />

envelope of CSR hypothesis. Higher values of APA coefficient of variations would indicate<br />

significant deviations from CSR towards aggregation.<br />

3. Result<br />

At first I have studied the relation between APA <strong>and</strong> the main biometrical attributes. There were<br />

identified very significant correlations with low intensity of APA with sapling diameter (r =<br />

0.23***) <strong>and</strong> crown diameter (r = 0.21***). This is an expected result because several<br />

researchers (Moore et al., 1973; Smith, 1987; Winsauer <strong>and</strong> Mattson, 1992) mentioned<br />

correlations of mature trees APA with the diameter or basal area.<br />

Obviously there is a strong negative correlation between APA <strong>and</strong> sapling density (r = - 0.99<br />

***) <strong>and</strong> even between APA coefficient of variation <strong>and</strong> density (r = - 0.72 *). The uniformity<br />

tendency is more evident at a higher density.<br />

Several studies (Moore et al., 1973; Winsauer <strong>and</strong> Mattson, 1992) indicate that APA might<br />

be correlated with growth <strong>and</strong> competition. Competition is one of the processes that shape the<br />

saplings spatial distribution. Consequently there were analyzed the correlations between APA<br />

<strong>and</strong> competition indices – Schutz index <strong>and</strong> Hegyi index computed in respect to diameter,<br />

height, crown volume <strong>and</strong> crown external surface. The strongest correlation is between APA<br />

<strong>and</strong> the Hegyi index calculated in respect to diameter (r = - 0.32 ***) <strong>and</strong> height (r = - 0.27 ***).<br />

In order to evaluate the performance of growth it was analyzed the correlation between APA<br />

<strong>and</strong> sapling height growth. Surprisingly, there is no correlation between these parameters (r =<br />

0.08 *). Other studies indicated significant correlations between mature trees growth (diameter<br />

growth) <strong>and</strong> APA but sapling populations seem to be more dynamic than mature trees. This<br />

initial developing stage is very unstable regarding spatial distribution of the individuals so APA<br />

is a factor with a smaller impact on height growth.<br />

Some authors (Mercier, Baujard, 1997) point out there is a relation between APA values or<br />

APA coefficient of variation <strong>and</strong> the spatial pattern of a population of mature trees. This fact<br />

seems to be relevant to sapling populations, because of the APA correlations with spatial pattern<br />

indicators – Morisita (r = 0.70 *) <strong>and</strong> Clark-Evans (r = -0.84 **). The APA coefficient of<br />

variation is also correlated with spatial pattern indicators Morisita (r = 0.88 **) <strong>and</strong> Clark-Evans<br />

(r = -0.58). Aggregated patterns lead to higher values of APA coefficient of variation (Figure 2).<br />

Figure 2: APA coefficient of variation values show aggregated patterns for both<br />

Morisita <strong>and</strong> Clark-Evans indices<br />

This detail shows that APA coefficient of variation might be used as an indicator of spatial<br />

distribution. The Monte-Carlo simulations indicate significant deviations from CSR towards<br />

aggregation - the values of APA coefficient of variation overcome the 95% confidence envelope<br />

for all the plots (Figure 3).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Palaghianu 2010. The use of Voronoi tessellation to characterize sapling populations<br />

83<br />

Figure 3: APA coefficient of variation values exceed the higher limits of 95% confidence envelope<br />

4. Discussion<br />

The results indicated that saplings APA have poor relationships with biometrical attributes. The<br />

reason might be the fact that APA takes into account only sapling position <strong>and</strong> no other<br />

biometrical feature. There is a solution - the generation of Voronoi weighted diagram in respect<br />

to some biometric parameter. VORONOI software has the capability to generate such diagrams.<br />

Still, there is one problem because it’s very difficult to compute the area of the resulted cells.<br />

In saplings population APA can be described as a low performance indicator of competition<br />

because there is no relation to height growth <strong>and</strong> there are low intensity correlations with<br />

competition indices. There might be a possibility to use APA in competition analyses in<br />

combination with other attributes, but not as a st<strong>and</strong>-alone indicator. However, one important<br />

aspect in assessing competition is that the non-weighted diagrams are the best mathematical<br />

solution to establish the neighbours of a sapling or tree. So APA might be used as a criterion for<br />

selecting neighbours. APA coefficient of variation is a straight-forward indicator with positive<br />

results as an indicator of spatial pattern. The significance of this indicator might be evaluated by<br />

comparing the results with the values of a confidence envelope.<br />

APA in sapling populations is a complex <strong>and</strong> useful tool for characterizing population structure<br />

regarding spatial distribution, but seems more suited to mature trees.<br />

I hope the development of software tools VORONOI <strong>and</strong> ARIA VORONOI will simplify <strong>and</strong><br />

support further studies.<br />

References<br />

Bella, I.E., 1971. A new competition model for individual trees, <strong>Forest</strong> Science, 17: 364–372.<br />

Brown, G.S., 1965, Point density in stems per acre, New Zeal<strong>and</strong> <strong>Forest</strong>ry Service Research<br />

Notes, 38: 1-11.<br />

Clark, P. J., Evans, F. C.1954. Distance to nearest neighbour as a measure of spatial<br />

relationships in populations. Journal of Ecology, 35: 445-453.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Palaghianu 2010. The use of Voronoi tessellation to characterize sapling populations<br />

84<br />

Garcia, O., 2008. Plant individual-based modelling: More than meets the eye, World<br />

Conference on Natural Resource Modeling Warsaw,<br />

http://forestgrowth.unbc.ca/warsaw.pdf.<br />

Hegyi, F., 1974. A simulation for managing jack-pine st<strong>and</strong>s, Growth Models for Tree <strong>and</strong><br />

St<strong>and</strong> Simulation. Royal College of <strong>Forest</strong>ry, Stockholm, Sweden, 74–90.<br />

Konig, G., 1835. Die Forst-Mathematik in den Grenzen wirtschaftlicher Anwendung nebst<br />

Hülfstafeln für die Forstschätzung und den täglichen Forstdienst, 4 th Ed - 1854, 830 p.<br />

Mead, R., 1966. A relationship between individual plant-spacing <strong>and</strong> yield. Annals of Botany<br />

30: 301-309.<br />

Mercier, F., Baujard, O., 1997. Voronoi diagrams to model forest dynamics in French Guiana,<br />

GeoComputation ‘97 & SIRC ‘97 Proceedings, University Otago, New Zeal<strong>and</strong>, 161-171.<br />

Moore, J.A., Budelsky. C.A., Schlesinger, R.C., 1973, A new index representing individual tree<br />

competitive status, Canadian Journal of <strong>Forest</strong> Research 3: 495-500.<br />

Morisita, M., 1962. Iδ index, a measure of dispersal of individuals. Researches on Population<br />

Ecology 4: 1-7.<br />

Ohyama, T., 2008. Voronoi diagram, http://www.nirarebakun.com/eng.html;<br />

Owens, M.K., Norton, B.E., 1989. The impact of ‘available area’ on Artemisia tridentata<br />

seedling dynamics. Vegetatio 82: 155-162.<br />

Perry, G.L.W., 2004. SpPack: spatial point pattern analysis in Excel using Visual Basic for<br />

Applications (VBA), Environmental Modelling & Software 19: 559–569.<br />

Schutz, J.P., 1989. Zum Problem der Konkurrenz in Mischbeständen. Schweiz. Z. Forstwes 140:<br />

1069–1083.<br />

Smith, W.R., 1987. Area potentially available to a tree: a research tool, The 19th Southern<br />

<strong>Forest</strong> Tree Improvement Conference, Texas, 29 p.<br />

Staebler, G.R., 1951. Growth <strong>and</strong> spacing in an even-aged st<strong>and</strong> of Douglas fir, Master’s thesis,<br />

University of Michigan, 46 p.<br />

Wichmann, L., 2002. Modelling the efects of competition between individual trees in forest<br />

st<strong>and</strong>s, PhD Thesis, The Royal Veterinary <strong>and</strong> Agricultural University, Copenhagen,<br />

112p.<br />

Winsauer, S.A., Mattson, J.A., 1992. Calculating Competition In Thinned Northern Hardwoods,<br />

Research Paper NC-306, St. Paul, USDA, 10 p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Peña-Cortés et al. 2010. Spatial <strong>and</strong> temporal dynamics <strong>and</strong> future trends of change in La Araucania, Chile<br />

85<br />

Spatial <strong>and</strong> temporal dynamics <strong>and</strong> future trends of change in the<br />

coastal l<strong>and</strong>scape of La Araucania, Chile<br />

Fern<strong>and</strong>o Peña-Cortés 1* , Daniel Rozas 1 , Enrique Hauenstein 1 , Carlos Bertrán 2 , Jaime<br />

Tapia 3 & Marco Cisternas 4<br />

1 Laboratorio de Planificación Territorial, Escuela de Ciencias Ambientales, Facultad de<br />

Recursos Naturales, Universidad Católica de Temuco, Chile<br />

2 Instituto de Zoología, Universidad Austral de Chile, Valdivia, Chile<br />

3 Instituto de Química de Recursos Naturales, Universidad de Talca, Chile<br />

4 Facultad de Recursos Naturales, Escuela de Ciencias del Mar, Pontificia Universidad<br />

Católica de Valparaíso, Chile<br />

Summary<br />

In the coastal rim of La Araucania l<strong>and</strong> use has significantly changed during the last 100 years<br />

mainly due to anthropic effects, though to significant natural events too. The aim of this study is<br />

to identify the direction <strong>and</strong> magnitude of change patterns that l<strong>and</strong>scape units have had from<br />

1980 to 2007. The analysis was made in base of information obtained by remote sensing <strong>and</strong><br />

cadastral mapping of vegetation, evaluating the change of spatial patterns to propose a<br />

prospective model to 2017. The prospective method was based on Markov chain <strong>and</strong> Cellular<br />

automata. The results show significant changes in usage patterns, denoting especially forest<br />

expansion <strong>and</strong> fragmentation of natural habitats. For the year 2017, the survey indicates a trend<br />

of maintaining the forest, although, however, is expected for a stabilization in the fragmentation<br />

process.<br />

Keywords: Coastal rim, l<strong>and</strong>scape units, time series, Markov chains.<br />

1. Introduction<br />

The change in l<strong>and</strong> use is one of the most relevant events in the relation between man <strong>and</strong><br />

environment. While in many cases the spatial configuration of the territory is due to natural<br />

events, are human activities, in particular, economic activity, one of the main agents of change<br />

modelers (Van der Veen & Otter 2001; Fan et al. 2008; Koomen et al. 2008).<br />

According to Peña-Cortés et al. (2009), original l<strong>and</strong> use in coastal watersheds of La Araucania<br />

has significantly changed in the last 100 years. L<strong>and</strong>scape has been strongly altered by man <strong>and</strong><br />

by important natural events like the earthquake <strong>and</strong> tsunami of 1960, <strong>and</strong> the recent one from<br />

February 2010, creating a cultural l<strong>and</strong>scape dominated by an agricultural matrix <strong>and</strong> a matrix<br />

exponential occupation of forest from 30 years ago. In this sense, temporal analysis of<br />

l<strong>and</strong>scape dynamics provides us with, on the one h<strong>and</strong>, some relevant information about<br />

ecological processes associated to magnitude <strong>and</strong> direction of the changes (Torrejón & Cisternas<br />

2002; Bender et al.2005; Peña et al. 2006), also a historical synthesis of socio-economic events<br />

which resulted in the territory, <strong>and</strong> the evolution of state policies concerning the use of natural<br />

resources.<br />

* Corresponding author. Tel.: 56-45-205469 - Fax: 56-45-205469<br />

Email address: fpena@uct.cl<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Peña-Cortés et al. 2010. Spatial <strong>and</strong> temporal dynamics <strong>and</strong> future trends of change in La Araucania, Chile<br />

86<br />

Numerous investigations have attempted to develop scientists <strong>and</strong> planners on the dynamics of<br />

change <strong>and</strong> development of future projections of l<strong>and</strong> occupation (Baker 1989; Cousins2001;<br />

Weng 2002; Luijten 2003; Jackson et al. 2004; Gómez-Mendoza et al. 2006; Fan et al. 2008;<br />

Wang et al. 2010), being widely used scenario <strong>and</strong> trend analysis (Koomen et al. 2008).<br />

This paper presents an application of a series of indices <strong>and</strong> metrics related to the structure <strong>and</strong><br />

temporary dynamics of l<strong>and</strong>scape in coastal watersheds of La Araucania. The objective is to<br />

identify the direction <strong>and</strong> magnitude of usage patterns that l<strong>and</strong>scape units have had from 1980<br />

to 2007. It also proposes a scenario to 2017 based on Markov Chain <strong>and</strong> Cellular Automata in a<br />

GIS environment.<br />

2. Methodology<br />

The study area (figure 1) corresponds to the coastal rim of La Araucania, located between 38°<br />

30’ y 39° 30’ of South Latitude <strong>and</strong> 72° 50’ y 73° 30’ of West Latitude. Its surface is 221.993<br />

hectares distributed in four coastal watersheds: Moncul (44.747 ha), Budi (48.494 ha), Chelle<br />

(9.267 ha) y Queule river (69.144 ha), which are on a territory formed by mountain ranges,<br />

marine erosion platforms <strong>and</strong> extensive fluvial <strong>and</strong> marine plains (Peña-Cortes et al. 2006).<br />

According to Di Castri <strong>and</strong> Hajek (1976) the weather is oceanic with a Mediterranean influence<br />

<strong>and</strong> an average annual rainfall from 1200 mm to 1600 mm.<br />

2.1 Mapping process<br />

For the identification <strong>and</strong> analysis of l<strong>and</strong> use coverages were used 1:60,000 scale aerial<br />

photographs for 1980, 1:20,000 For 1994 <strong>and</strong> also the regional classification of the cadastre<br />

vegetation resources of Chile (CONAF - CONAMA- BIRF 1999) updated in 2007 in 1:50.000<br />

scale. In all cases we used satellite imagery in support of L<strong>and</strong>sat <strong>and</strong> Aster sensor.<br />

On these images we proceeded to classify the various ground covers in a previously established<br />

categorization of 14 classes. For the analysis of spatial patterns <strong>and</strong> dynamic changes of use<br />

layers were reclassified into eight classes according to the type of coverage. The processing <strong>and</strong><br />

data analysis was performed with the software Taiga Idrisi <strong>and</strong> ArcGis 9.3.<br />

2.2 Analysis of l<strong>and</strong>scape<br />

The analysis of dynamics <strong>and</strong> patterns of l<strong>and</strong>scape considered calculating magnitude <strong>and</strong><br />

direction of changes during the assessed process. Among the first one, rates of change annual<br />

average for each class was obtained (TCC). For the detailed study of its temporal variation<br />

metrics were used at l<strong>and</strong>scape related to the fragments, edges <strong>and</strong> complex forms described by<br />

Patton (1985) <strong>and</strong> Henao (1988). The heterogeinity of the l<strong>and</strong>scape mosaic was evaluated by<br />

the diversity index <strong>and</strong> Shannon evenness. Regarding the direction of change, it was obtained<br />

by an evaluation pixel by pixel to its original state at time 0 to the next state at time 1. The result<br />

is given by the pixel number of class x (1-8) are transformed to the class <strong>and</strong> (1-8) over a period<br />

of time t (0-1).<br />

2.3 Prospective model<br />

For the development of a future projection for the current dynamics of the l<strong>and</strong>scape, the<br />

method of Markov Chain was used. This produces a transition probability matrix for each<br />

l<strong>and</strong>scape unit simulating a future scenario from the two previous statements. After the<br />

application of Markov chain, a stochastic projection algorithm was applied to evaluate the<br />

probability of each pixel to belong to either category of l<strong>and</strong>scape. Finally, to address the lack of<br />

spatial dependence <strong>and</strong> topological criteria in evaluating the future scenario, we employed a<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Peña-Cortés et al. 2010. Spatial <strong>and</strong> temporal dynamics <strong>and</strong> future trends of change in La Araucania, Chile<br />

87<br />

Cellular Automata algorithm, increasing the likelihood of belonging to a category based on its<br />

local proximity.<br />

3. Results<br />

During the period under review has seen a significant increase in the surface of the forest matrix<br />

dominated over other ground coverage, especially during the first 14 years of analysis.<br />

Moreover, the transition matrix classes <strong>and</strong> beaches <strong>and</strong> dunes, have experienced a steady<br />

decline over time. Regarding the projection to 2017 shows that the trend will continue with an<br />

increase forestry positive as the increase in the native forest, although the latter to a lesser<br />

extent. Table 1 shows the annual change rates for each type of l<strong>and</strong>scape.<br />

According to metrics that have characterized the evolution of l<strong>and</strong>scape, it is observed a trend to<br />

increase the number of fragments to the end of the period of analysis, which however, slightly<br />

decreases in 2017 scenario. While 1994 showed a lower number of fragments, their average size<br />

was the largest of the series, which was associated with a low density of edges. The average<br />

shape index indicates that the fragments are generally amorphous except 2017 which<br />

corresponds to oblong oval, also presented an average fractal dimension gives them a moderate<br />

complexity. In relation to the indices of diversity were high for all years with patterns of<br />

evenness than 60% in all cases. The metrics of each year can be seen in Table 2.<br />

Finally, in table 3 is presented the direction of changes occurred in l<strong>and</strong>scape units beginning in<br />

1980 represented by rows, through its new state in 2007 represented by columns. Units with<br />

more significant changes were wetl<strong>and</strong>s <strong>and</strong> other wet l<strong>and</strong>s, which became part of the<br />

agricultural matrix with almost 15.000 hectares.<br />

4. Discussion<br />

L<strong>and</strong>scape from the coastal rim of La Araucania is characterized by a marked anthropic<br />

difference (Peña et al. 2006, 2009). A signal of this is that prevailed coverages are related to<br />

productive activities like agriculture, silviculture <strong>and</strong> developing of human settlements, being<br />

these the one which presented major increases in surface from 1980 to 2007.<br />

Regarding l<strong>and</strong>scape patterns, showed a notable increase in the number of patches during the<br />

first 27 years of study, accompanied by a higher density <strong>and</strong> lower edges of medium size, what<br />

constitutes clear indicators of processes of fragmentation <strong>and</strong> habitat loss <strong>and</strong> it is also<br />

accentuated by the irregular shapes. In relation to the future scenario, it shows a tendency<br />

towards stabilization in the above processes, expressed mainly in a decrease in the number of<br />

fragments, the density of edges <strong>and</strong> shapes more regular <strong>and</strong> less complex than previous ones.<br />

Finally, in the period 1980-2007 the direction of changes greatly affected the natural areas with<br />

native forests, wetl<strong>and</strong>s <strong>and</strong> thicket, becoming part of the forestry <strong>and</strong> agricultural matrix,<br />

however, the latter also decreases in surface transition to forest matrix, which was identified as<br />

the main agent of change in the current configuration of the l<strong>and</strong>scape of the coast rim.<br />

References<br />

Baker, W., 1989., A review of models of l<strong>and</strong>scape change. L<strong>and</strong>scape Ecology, 2: 111-133.<br />

Bender, O., Boehmer, H.J., Jens, D <strong>and</strong> Schumacher, K.P., 2005. Using GIS to Analyze Long-<br />

Term cultural L<strong>and</strong>scape change in Southern Germany. L<strong>and</strong>scape Urban Plann,<br />

70:111-125.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Peña-Cortés et al. 2010. Spatial <strong>and</strong> temporal dynamics <strong>and</strong> future trends of change in La Araucania, Chile<br />

88<br />

Cousins, S., 2001., Analysis of l<strong>and</strong>-cover transitions based on 17th <strong>and</strong> 18th century cadastral<br />

maps <strong>and</strong> aerial photographs. L<strong>and</strong>scape Ecology, 16: 41–54.<br />

CONAF-CONAMA-BIRF., 1999. Catastro y evaluación de recursos vegetacionales nativos de<br />

Chile. Informe regional IX Región, Proyecto CONAF, CONAMA, BIRF, Santiago,<br />

Chile, 90 p.<br />

Di Castri, F. <strong>and</strong> Hajek, E., 1976. Bioclimatología de Chile. Ediciones Universidad Católica de<br />

Chile, Santiago, 128p.<br />

Fan, F., Wang, Y. <strong>and</strong> Wang, Z., 2008. Temporal <strong>and</strong> spatial change detecting (1998–2003) <strong>and</strong><br />

predicting of l<strong>and</strong> use <strong>and</strong> l<strong>and</strong> cover in Core corridor of Pearl River Delta (China) by<br />

using TM <strong>and</strong> ETM+ images. Environmental Monitoring Assessment, 137: 127-147.<br />

Gómez-Mendoza, L.,Vega-Peña, E., Ramírez, M., Palacio-Prieto, J.L. <strong>and</strong> Galicia, L., 2006.<br />

Projecting l<strong>and</strong>-use change processes in the Sierra Norte of Oaxaca, Mexico. Applied<br />

Geography, 26: 276–290.<br />

Henao, S., 1988. Introducción al manejo de cuencas hidrográficas. Universidad de Santo<br />

Tomas, Centro de Enseñanza Desescolarizada, Ediciones Usta, Bogotá, Colombia, 398<br />

p.<br />

Jackson, L., Bird, S., Matheny, R., O’neill, R., White, D., Boesch, K. <strong>and</strong> Koviach, J., 2004. A<br />

regional approach to projecting l<strong>and</strong>-use change <strong>and</strong> resulting ecological vulnerability.<br />

Environmental Monitoring <strong>and</strong> Assessment, 94: 231–248.<br />

Koomen, E., Rietveld, P. <strong>and</strong> de Nijs, T., 2008. Modelling l<strong>and</strong>-use change for spatial planning<br />

support. The Annals of Regional Science, 42: 1-10.<br />

Luijten, J.C., 2003. A systematic method for generating l<strong>and</strong> use patterns using stochastic rules<br />

<strong>and</strong> basic l<strong>and</strong>scape characteristics: results for a Colombian hillside watershed.<br />

Agriculture, Ecosystems <strong>and</strong> Environment, 95: 427–441<br />

Patton D.R., 1975. A diversity index for quantifying habitat edge. Wildlife Society Bulletin, 394:<br />

171-173.<br />

Peña-Cortés, F., Rebolledo, G., Hermosilla, K., Hauenstein, E., Bertrán, C., Schlatter, R. <strong>and</strong><br />

Tapia, J., 2006. Dinámica del paisaje para el período 1980-2004 en la cuenca costera del<br />

Lago Budi, Chile. Consideraciones para la conservación de sus humedales. Ecología<br />

Austral, 16:183-196.<br />

Peña-Cortés, F., Escalona-Ulloa, M., Rebolledo, G., Pincheira-Ulbrich, J. <strong>and</strong> Torres-Alvarez,<br />

O., 2009. Efecto del cambio en el uso del suelo en la economía local: una perspectiva<br />

histórica en el borde costero de La Araucanía, sur de Chile. En: U. Confalonieri, M.<br />

Mendoza, <strong>and</strong> L. Fernández (Eds.). Efecto de los cambios globales sobre la salud<br />

humana y la seguridad alimentaria. Buenos Aires, Argentina: RED CYTED: 184-197<br />

406RT0285.<br />

Torrejón, F. <strong>and</strong> Cisternas, M., 2002. Alteraciones del paisaje ecológico araucano por la<br />

asimilación mapuche de la agroganadería hispano-mediterránea (siglos XVI y XVII).<br />

Revista Chilena de Historia Natural,75(4):729-736.<br />

van der Veen, A. <strong>and</strong> Otter, H.S., 2001. L<strong>and</strong> use changes in regional economic theory.<br />

Environmental Modeling <strong>and</strong> Assessment, 6: 145-150.<br />

Weng, Q., 2002. L<strong>and</strong> use change analysis in the Zhujiang Delta of China using satellite remote<br />

sensing, GIS <strong>and</strong> stochastic modeling. Journal of Environmental Management, 64: 273–<br />

284.<br />

Wang, S.Y., Liu, J.C. <strong>and</strong> Ma, T.B., 2010. Dynamics <strong>and</strong> changes in spatial patterns of l<strong>and</strong> use<br />

in Yellow River Basin, China. L<strong>and</strong> Use Policy, 27: 313–323.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Peña-Cortés et al. 2010. Spatial <strong>and</strong> temporal dynamics <strong>and</strong> future trends of change in La Araucania, Chile<br />

89<br />

Table 1: Variation between years for types of l<strong>and</strong>scapes present in the coastal rim of La Araucanía.<br />

L<strong>and</strong>scape units 1980-1994 (%) 1994-2007 (%) 2007-2017 (%)<br />

Human settlement 4,00 1,30 -0,50<br />

Native forest -1,90 0,30 0,30<br />

Waterbody 0,80 0,00 -0,90<br />

Wetl<strong>and</strong> <strong>and</strong> other wet terrains -10,20 2,10 -2,30<br />

Agricultural matrix 3,00 -1,50 -0,70<br />

Transition matrix -8,40 -2,50 -1,00<br />

<strong>Forest</strong>ry matrix 15,70 5,60 2,20<br />

Beaches <strong>and</strong> dunes -0,40 -5,00 -3,30<br />

Table 2: Temporal variation of the metrics of l<strong>and</strong>scape in coastal rim of La Araucania<br />

TA =total area; NP= total patches; MPS= medium patches size; ED= edges density; TE= total edges;<br />

medium shape index; MFRACT= medium fractal dimension.<br />

Metrics 1980 1994 2007 2017<br />

TA (Ha) 166.034 165.835 165.835 165.831<br />

NP 911 873 2.221 1.763<br />

MPS 166.38 288.24 167.58 105.08<br />

ED (m/Ha) 60.03 46.07 74.94 50.87<br />

TE (Km) 9.967 7.64 12.427 8.435<br />

MSI 2,46 2,02 2,1 1,65<br />

MFRACT 1,32 1,3 1,37 1,29<br />

Shannon's Diversity 1,26 1,36 1,46 1,45<br />

Shannon's Evenness 0,75 0,65 0,7 0,69<br />

Dominance 0,51 0,71 0,61 0,62<br />

Table 3: Direction of change in l<strong>and</strong>scape units in the coastal rim of La Araucania.<br />

1= human settlements; 2= native forest; 3=waterbody; 4= Wetl<strong>and</strong> <strong>and</strong> other wet terrains; 5= agricultural<br />

matrix; 6= transition matrix; 7= <strong>Forest</strong>ry matrix; 8= beaches <strong>and</strong> dunes<br />

2007<br />

1 9 8 0<br />

L<strong>and</strong>scape<br />

Units 1 2 3 4 5 6 7 8<br />

1 146 0 5 0 8 0 1 2<br />

2 3 14.343 38 952 8.567 1.470 8.475 25<br />

3 9 45 7.018 233 557 46 33 71<br />

4 105 2.509 1.343 6.035 13.908 591 2.265 115<br />

5 65 3.935 481 974 48.876 1.031 7.921 12<br />

6 0 4.684 6 166 6.207 2.798 14.780 18<br />

7 0 367 0 16 311 23 1.108 1<br />

8 14 4 115 98 600 433 552 1.318<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Peña-Cortés et al. 2010. Spatial <strong>and</strong> temporal dynamics <strong>and</strong> future trends of change in La Araucania, Chile<br />

90<br />

Figure 1: Study area. Elaborated by author.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F.M. Rabenilalana et al. 2010. Multi-temporal analysis of forest l<strong>and</strong>scape fragmentation in the North East of Madagascar<br />

91<br />

Multi-temporal analysis of forest l<strong>and</strong>scape fragmentation in the North<br />

East of Madagascar<br />

F.M. Rabenilalana 1 , L.G. Rajoelison 1 , J.P. Sorg 2 , J.L. Pfund 3 & H Rakoto Ratsimba 1<br />

1 Département des Eaux et Forêts, Ecole Supérieure des Sciences Agronomiques,<br />

Université d’Antananarivo, Madagascar<br />

2 Swiss Federal Institute of Technology Zurich, Institut of Terrestrial Ecosystems,<br />

Groupe de foresterie pour le développement, Zurich. Switzerl<strong>and</strong><br />

3 CIFOR, Center for International <strong>Forest</strong>ry Research, Bogor, Indonesia<br />

Abstract<br />

Habitat fragmentation caused by insufficient arable l<strong>and</strong> <strong>and</strong> rural poverty has become a main<br />

concern to tropical forest managers in Madagascar. In order to inform management decisions,<br />

the present study aims to provide quantitative information on the fragmentation process <strong>and</strong> its<br />

driving forces in Manompana. Therefore, time series analysis of fragmentation in relation to the<br />

spatial distribution of settlements, roads <strong>and</strong> topographic parameters was done. Results showed<br />

that (1) deforestation has been increasing due to exp<strong>and</strong>ing upl<strong>and</strong> rice production, thus<br />

progressively increasing l<strong>and</strong>scape fragmentation (2) fragmentation dynamic significantly<br />

differed with distance to paths, villages <strong>and</strong> depending on topographic characteristics. In the<br />

following these parameters will be used for spatio-temporal modeling of the l<strong>and</strong>scape<br />

dynamics to support decision making on sustainable management of natural resources.<br />

Keywords: l<strong>and</strong>scape, fragmentation, tropical humid forest, spatial analysis, Madagascar.<br />

1. Introduction<br />

Fragmentation refers to breaking a whole into smaller pieces while controlling for changes in<br />

the amount of habitat (Forman 1995, Fahrig 2003, Ewers 2006). Deforestation is one driver of<br />

habitat fragmentation. Fragmentation often results in the loss of habitat <strong>and</strong> the isolation of the<br />

remaining forest fragments. It creates matrices of human-managed areas, secondary vegetation<br />

regrowth, <strong>and</strong> fragments of primary forest (Benitez-Malvido 2003).<br />

The region of Manompana, is a disturbed tropical lowl<strong>and</strong> rainforest,forming a forest corridor<br />

between two protected areas with very high levels of endemism: the special reserve<br />

Ambatovaky in the South West <strong>and</strong> Biosphere Reserve of Mananara Nord in the North. In the<br />

region of Manompana forest clearing due to slash <strong>and</strong> burn agriculture <strong>and</strong> commercial logging<br />

are the main drivers of deforestation <strong>and</strong> thus habitat fragmentation, affecting the l<strong>and</strong>scape<br />

structure, the livelihoods <strong>and</strong> the biodiversity (Forman & Godron 1986, Andren 1992, Reed<br />

1996, Franklin 2001, McGarigal 2002, Kuppfer 2006, Jaeger 2000, Collinge 2009).<br />

This paper aims to study spatio-temporal patterns of l<strong>and</strong>scape fragmentation at a regional scale,<br />

i.e. the fragmentation of the forest ecosystem in Manompana in order to inform management<br />

decisions.<br />

Therefore, l<strong>and</strong> cover predictions for future were made using SPOT 5 multispectral, multib<strong>and</strong><br />

images (V 0,50 -0,59 μm R 0,61 -0,68 μm Near InfraRed or NIR 0,78 -0,89 μm <strong>and</strong> Mean<br />

InfraRed or MIR 1,58 -1,75 μm) with a 10 meters resolution from 2004 <strong>and</strong> 2009. Furthermore,<br />

1 Corresponding author. Tel.:+2613367162<br />

Email address: rmihajamanana@yahoo<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F.M. Rabenilalana et al. 2010. Multi-temporal analysis of forest l<strong>and</strong>scape fragmentation in the North East of Madagascar<br />

92<br />

we focused on identifying the main factors that play a direct role in shaping the spatial<br />

variations in forest cover.<br />

We will start with a description of the data sources <strong>and</strong> the process of analysis adopted in the<br />

study <strong>and</strong> finally discuss directions <strong>and</strong> needs regarding the sustainable management of natural<br />

resources.<br />

2. Methodology<br />

2.1 Study area<br />

Manompana is situated on the eastern coast of Madagascar (752144 in the East <strong>and</strong> 1041646 in<br />

the north, Laborde Coordinates). This region experiences a warm <strong>and</strong> humid climate. Rainfall is<br />

reliable, <strong>and</strong> arises throughout the year with a mean annual rainfall of 3677 mm (Weather<br />

station at Soanieràna Ivongo, 2009). The wettest months are March <strong>and</strong> April at the beginning<br />

of winter. The temperature lies always above 21°C with a mean annual of 23,7°C (Weather<br />

station at Soanieràna Ivongo, 2009). The forest vegetation is typical of tropical lowl<strong>and</strong><br />

rainforest dominated by Anthostema madagascariensis <strong>and</strong> family of Myristicaceae (Guillaumet<br />

<strong>and</strong> Humbert 1975).<br />

2.1.2 Data processing <strong>and</strong> classification<br />

There are many methods for detecting l<strong>and</strong> cover changes available in the literature such as<br />

image differencing <strong>and</strong> post-classification. In the current study, image differencing was adopted<br />

(Singh 2009). This method is used mainly to detect areas with significant l<strong>and</strong> cover changes<br />

<strong>and</strong> does not require atmospheric correction, but requires careful classification of “change /<br />

persistence” thresholds.<br />

In detail we followed the succeeding steps. First, the study area was extracted from the<br />

respective scene. In a second step, the forest components were identified using the classic<br />

method of radiometric intervals. The unsupervised classification was done using the application<br />

ISODATA (Iterative Self-Organizing Data Analysis Technics), the supervised classification<br />

was done using the application Maximum Likelihood in Arc Map 9.2, Arc view 3.2a <strong>and</strong> ENVI<br />

or Environment for Visualizing Images software (Gonzales 1988, Barrett <strong>and</strong> Curtis 1976).<br />

First, ISODATA calculated class means pixels evenly distributed in the data space.<br />

Consequently, it iteratively clustered the remaining pixels according to their reflectance<br />

(including the diffuse radiation by the atmosphere, the radiation reflected by the pixel <strong>and</strong> the<br />

radiation reflected from neighbouring pixels) using minimum distance techniques. Each<br />

iteration recalculated the means <strong>and</strong> reclassified pixels with respect to the new (class) means.<br />

This process continued until the number of pixels in each class changed by less than the selected<br />

pixel change threshold or until the maximum number of iterations is reached. Next, Maximum<br />

Likelihood methods were applied. This supervised classification assumed that the values for<br />

each class in each b<strong>and</strong> are normally distributed <strong>and</strong> calculated the probability that a given pixel<br />

belongs to a specific class. Unless a probability threshold was selected, all pixels were classified.<br />

Each pixel was assigned to the class (forest, non forest) that had the highest probability.<br />

Finally, the fragmentation index was analyzed by pattern analysis. Fragmentation is a measure<br />

of the ecological quality of a habitat. In order to compute it, the following kernel-based formula<br />

(1) was used:<br />

Fragmentation = (n - l)/(c - 1) (1)<br />

where n is the number of different l<strong>and</strong>-cover classes present in a kernel, <strong>and</strong> c the number of<br />

cells considered (Monmonier, 1974). The kernel represents the similarity between two cells<br />

defined as dot-product in the new vector space.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F.M. Rabenilalana et al. 2010. Multi-temporal analysis of forest l<strong>and</strong>scape fragmentation in the North East of Madagascar<br />

93<br />

For this computation, a 3 by 3 kernel (majority filter) was used so that c equalled 9.This interval<br />

choice was linked to the importance of reflectance. A reflectance of 100% means that the<br />

surface reflects all solar energy in the atmosphere. This corresponds to a fragmentation index of<br />

0. When the reflectance is 0%, the fragmentation index equals 1.<br />

2.1.3 <strong>Change</strong> analysis<br />

L<strong>and</strong> cover maps from two different dates, 2004 <strong>and</strong> 2009, were used to make l<strong>and</strong> cover<br />

predictions for 2010. Using L<strong>and</strong> <strong>Change</strong> Modeler (Eastman 2006), this was done in two major<br />

stages: the transition potential sub-model, which is developed in this paper, stage <strong>and</strong> the<br />

change prediction model stage. A transition sub-model consists of a single l<strong>and</strong> cover transition<br />

or a group of transitions that are thought to have the same underlying driver variables. In the<br />

first stage, particular transitions of interest for the sub-model <strong>and</strong> the variables which are<br />

assumed to drive the transitions taking place were specified. Variables considered as drivers of<br />

deforestation were distance to rivers <strong>and</strong> streams, paths, villages <strong>and</strong> to topographic parameters.<br />

Using Cramer’s V or Phi statistic the strength, association or dependency between two variables<br />

(Agresti 2002) was analyzed. The closer V is to 0 when the smaller ie the association between<br />

the categorical variables X <strong>and</strong> Y. On the other h<strong>and</strong>, V being close to 1 is an indication of a<br />

strong association between two variables.<br />

Afterward, gain <strong>and</strong> loss of fragmentation (forests cover) area were calculated between two<br />

periods with the aim to determine the rate of fragmentation in the study area.<br />

3. Results<br />

3.1 Fragmentation Index<br />

The fragmentation was classified in two groups depending on the fragmentation index: a) a<br />

forest fragment was considered to have a low level of fragmentation when the index was<br />

between [0; 0.1[, b) a forest fragment was considered to have a high level of fragmentation<br />

when the index was within the interval [0.1; 1].<br />

Table 1 shows the fragmentation index. In 2004, the index ranged from 0 to 0.875 while in 2009<br />

the values were between 0 <strong>and</strong> 0.25. The reasons for this difference could be either a) the<br />

luminosity at the time when the images were taken was not the same or b) the forest fragments<br />

of 2004 have been converted into less fragments in 2009, which is not evident in the short<br />

period. In terms of patch numbers, there is a large difference between 2004 <strong>and</strong> 2009. 17 % of<br />

forests cover with low fragmentation in 2004 shows a high level of fragmentation or has been<br />

converted into other l<strong>and</strong> uses in 2009.<br />

Table 1: Fragmentation index characteristics<br />

Period<br />

Total<br />

area<br />

Count Minimum Maximum Mean<br />

St<strong>and</strong>ard<br />

deviation<br />

Low<br />

fragmented<br />

class<br />

Patches number<br />

2004 30487 18918482 0 0.875 0.054 0.01 65535 5633<br />

2009 22472 18918482 0 0.25 0.017 0.01 11231 6442<br />

Fragmented<br />

class<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F.M. Rabenilalana et al. 2010. Multi-temporal analysis of forest l<strong>and</strong>scape fragmentation in the North East of Madagascar<br />

94<br />

3.2 Fragmentation forest area<br />

In terms of forest area, a decrease (occurred during the period under review) was observed<br />

between the two periods. A deforestation rate of 5.25 % per year was calculated. Figure 1<br />

explains that 46% of low fragmented forest was changed into the other l<strong>and</strong> use classes in 2009.<br />

Approximately, most of the forested area lost has been converted into cropl<strong>and</strong> <strong>and</strong> only 36% in<br />

highly fragmented forest. In fact, there has been a tendency of highlighting small-scale<br />

migratory farmers <strong>and</strong> "poverty" as the major cause of l<strong>and</strong>scape change.<br />

In spite of increased fragmentation, the “gains <strong>and</strong> losses’ analysis showed that some fragments<br />

persisted. In reality, more or less 35% of the total forest area resists change. These values<br />

illustrate l<strong>and</strong>scape changes at a very high rate. And if there is no action taken soon, large<br />

amounts of habitat <strong>and</strong> biodiversity values will be lost.<br />

Area (Ha)<br />

35 000,00<br />

30 000,00<br />

25 000,00<br />

20 000,00<br />

15 000,00<br />

10 000,00<br />

5 000,00<br />

0,00<br />

2004 2009<br />

Year<br />

Low fragmented Fragmented<br />

Figure 1: <strong>Forest</strong> area change from 2004 to 2009<br />

3.2 <strong>Change</strong> analysis<br />

The impact of distance variables (distance to streams, rivers, paths, villages <strong>and</strong> topographic<br />

parameters) on l<strong>and</strong>scape changes was calculated. Cramer’s V test revealed that a) the selected<br />

distance variables did not influence low fragmented forest patches because V equal 0; b)<br />

distance to path <strong>and</strong> topographic parameters did, however, impact high fragmented forest<br />

patches (see figure 2).<br />

V*<br />

0,45<br />

0,4<br />

0,35<br />

0,3<br />

0,25<br />

0,2<br />

0,15<br />

0,1<br />

0,05<br />

0<br />

Village distance to<br />

fragmented forest<br />

Path distance to<br />

fragmented forest<br />

Stream & river distance<br />

to fragmented forest<br />

Topography to<br />

fragmented forest<br />

Variables distance<br />

Cramer's V* ( V significant)<br />

Figure 2: Test of driver variables<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F.M. Rabenilalana et al. 2010. Multi-temporal analysis of forest l<strong>and</strong>scape fragmentation in the North East of Madagascar<br />

95<br />

4. Discussion<br />

The annual deforestation rate was estimated to be 5.25%. This is very high in comparison to the<br />

national rate of 0.3 % (FAO, 2009). This means that the pressure on the forest in the area of<br />

Manompana is extremely high. In this part of the country, the local population does not produce<br />

enough rice to feed their family, there is a chronic shortage. The rice needs are merely satisfied<br />

on an average of nine months per year. The rest of the year people resort to other products,<br />

mainly cassava, yams <strong>and</strong> bananas, as it is often not possible to buy rice during this starvation<br />

period. The development of irrigated rice production is hampered by various factors, such as<br />

topographical, technical, socio economic <strong>and</strong> policy. Thus, to maximize production, slash <strong>and</strong><br />

burn agriculture has remained the main method of food production so far. Therefore, the<br />

deforestation is continuously increasing for upl<strong>and</strong> rice production <strong>and</strong> the l<strong>and</strong>scape is<br />

becoming progressively more fragmented. The l<strong>and</strong> use change map <strong>and</strong> examination of the<br />

driving variables of deforestation indicate that distance to topographic parameters <strong>and</strong> to paths<br />

has the highest impact on forest fragmentation, followed by distance to villages, rivers <strong>and</strong><br />

streams. Topography is an important factor impacting fragmentation. Most likely, because local<br />

people’s choice of l<strong>and</strong> clearing is strongly related to it. For example, local people prefer areas<br />

where slopes are not too steep <strong>and</strong> l<strong>and</strong> less windy because of climate. Paths are the only way of<br />

communication in inaccessible areas. Thus, to access a forestl<strong>and</strong>, the local population opened<br />

paths along the forest edge or in the forest itself.<br />

The result from this study will be able to contribute to the objectives of conservation. Indeed, if<br />

the variables remain more or less the same in the coming years, meaning that there is no<br />

opening of new paths or migration (village), <strong>and</strong> pertinence area doesn’t change, so biodiversity<br />

richness biodiversity is maintained.<br />

References<br />

Achard, F., Eva, H., Stibig, H.J., Mayaux, P., Gallego, T., Richards, J.P., Malingreau, J.P.,<br />

2002. Determination of deforestation rates of the world’s humid tropical forests. Science,<br />

297: 999-1002.<br />

Agresti, A., 2002. Categorical Data Analysis, Second Edition. Wiley, New York, USA. 734 pp.<br />

Andren, H., 1992. Corvid density <strong>and</strong> nest predation in relation to forest fragmentation: a<br />

l<strong>and</strong>scape perspective. Ecology, 73 (3):<br />

Benitez-Malvido, J. <strong>and</strong> Martinez-Ramos, M., 2003. Impact of <strong>Forest</strong> Fragmentation on<br />

Understory Plant Species Richness in Amazonia. Conservation Biology, 17: 389-400.<br />

Collinge, S. K., 2009. Ecology of Fragmented <strong>L<strong>and</strong>scapes</strong>. Baltimore: Johns Hopkins<br />

University Press.<br />

Eastman, J. R., 2006. IDRISI Andes, Tutorial. Clark University. IDRISI, production, Worcester,<br />

USA. 284pp.<br />

Ewers, R. <strong>and</strong> Didham, R. K., 2006. Confounding factors in the detection of species responses<br />

to habitat fragmentation. Biological Reviews, 18: 117-142.<br />

Fahrig, L., 2003. Effects of habitat fragmentation on Biodiversity. Annual Review of Ecology,<br />

Evolution <strong>and</strong> Systematics, 34: 487-515.<br />

FAO, 2009. Situation des forêts du monde. Organisation des nations Unies pour l’Agriculture et<br />

l’Agriculture. Viale delle Terme di Caracalla - 00153 Rome, Italie. 168pp.<br />

Forman, R. T. T., <strong>and</strong> Godron, M., 1986. L<strong>and</strong>scape ecology. John Wiley & Sons, Nex York.<br />

620pp.<br />

Forman, R. T. T., 1995. L<strong>and</strong> mosaics: The ecology of l<strong>and</strong>scapes <strong>and</strong> regions. Cambridge<br />

University Press, Cambridge, UK.<br />

Franklin, S., Stenhoise, G. B., Hansen, M., Popplewell, C. C., Dechka, J.A., Peddle, D.R., 2001.<br />

An integrated decision tree approach (IDTA) to mapping l<strong>and</strong>cover using satellite remote<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F.M. Rabenilalana et al. 2010. Multi-temporal analysis of forest l<strong>and</strong>scape fragmentation in the North East of Madagascar<br />

96<br />

sensing in support of grizzly bear habitat analysis in the alberta yellowhead ecosystem:<br />

remote sensing <strong>and</strong> spatial data integration: measuring, monitoring <strong>and</strong> modelling.<br />

Canadian Remote sensing Society, 27 (6):579-592.<br />

Guillaumet, J-L. <strong>and</strong> Koechlin, J., 1971. Contribution à la définition des types de végétation<br />

dans les régions tropicales (exemple de Madagascar). C<strong>and</strong>ollea, 26(2): 263-277.<br />

Jaeger, J. A. G., 2000. L<strong>and</strong>scape division, splitting index, <strong>and</strong> effective mes size: new<br />

measures of l<strong>and</strong>scape fragmentation. L<strong>and</strong>scape ecology, 15: 115-130.<br />

Kuppfer, J. A., Malanson, G.P., Franklin, S.B., 2006. Not seeing the ocean for the isl<strong>and</strong>s: the<br />

mediating influence of matrix-based processes on forest fragmentation effects. <strong>Global</strong><br />

Ecology <strong>and</strong> Biogeography, 15: 8-20.<br />

Monmoniern, M. S., 1974. Measures of Pattern Complexity for Choroplethic Maps.<br />

Cartography <strong>and</strong> Geographic Information Science, 1(2): 159-169.<br />

McGarigal, K. <strong>and</strong> Cushman, S.A., 2002.Comparative evaluation of experimental approaches to<br />

the study of habitat fragmentation effects. Ecological applications, 12: 335-345.<br />

Reed, J. M., 1996. Using statistical probability to increase confidence of inferrinf species<br />

extinction. Conservation Biology, 10: 1283-1285.<br />

Singh, A. 2009. Review Article Digital change detection technique using remotelysensed<br />

data. International Journal of Remote Sensing, 10(6): 989-1003.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ruskule et al. 2010. Patterns of afforestation process in ab<strong>and</strong>oned agriculture l<strong>and</strong> in Latvia<br />

97<br />

Patterns of afforestation process in ab<strong>and</strong>oned agriculture l<strong>and</strong> in<br />

Latvia<br />

Anda Ruskule * , Olģerts Nikodemus, Zane Kasparinska & Raimonds Kasparinskis<br />

University of Latvia, Latvia<br />

Abstract<br />

Afforestation of the former agriculture l<strong>and</strong> is one of the most typical trends of the<br />

contemporary Latvian rural l<strong>and</strong>scape. The study examines development of l<strong>and</strong>scape<br />

ecological succession, its spatial character <strong>and</strong> influencing environmental factors within<br />

ab<strong>and</strong>oned agriculture l<strong>and</strong> in Vidzeme, central part of Latvia. The study area comprises<br />

a great variety in the spatial dynamics of the ecological succession. Character of the<br />

ecological succession is influenced by various factors like soil fertility, moisture <strong>and</strong><br />

light conditions, species composition of surrounding forest, former l<strong>and</strong> use etc. The<br />

study aims to assess impacts of these factors on dynamics of afforestation process in<br />

order to predict its future development.<br />

Keywords: ab<strong>and</strong>onment, afforestation, l<strong>and</strong>scape ecological succession, environmental factors<br />

1. Introduction<br />

L<strong>and</strong> ab<strong>and</strong>onment <strong>and</strong> related natural afforestation process is typical feature of contemporary<br />

l<strong>and</strong>scape in marginal areas all over the Europe <strong>and</strong> most probably this trend will last for the<br />

coming years - modelling of l<strong>and</strong> use development in Europe until 2030 shows decline of<br />

agricultural l<strong>and</strong> as one of the major factors influencing European l<strong>and</strong>scape in future (Stoate et<br />

al 2009). In Latvia afforestation process has been noted from the beginning of the last century,<br />

resulting from frequentative change of political <strong>and</strong> economic situation – if in 1935 forest<br />

occupied ca. 27 % of the l<strong>and</strong> area, than by 2008 its area has almost doubled, reaching ca. 50 %<br />

(Ministry of Agriculture 2009; Nikodemus et al. 2005). During the last ten years forest area has<br />

increased mostly due to natural afforestation, twice exceeding artificial afforestation. This is<br />

related to extensive l<strong>and</strong> ab<strong>and</strong>onment after regaining of independence from Soviet Union in<br />

1991 <strong>and</strong> the following collapse of the collective farming system, urbanisation process <strong>and</strong> l<strong>and</strong><br />

reform (big share of the agriculture l<strong>and</strong> was retrieved by the former l<strong>and</strong> owners, presently<br />

living in city <strong>and</strong> having not much interest in farming).<br />

In the recent years l<strong>and</strong> ab<strong>and</strong>onment has become an important issue in scientific research,<br />

questioning what are the factors influencing this process, how it impacts the l<strong>and</strong>scape structure<br />

<strong>and</strong> biodiversity as well as the potential of ab<strong>and</strong>oned l<strong>and</strong> for use of natural resources or<br />

restoration of habitats. Socio-economic <strong>and</strong> political factors driving the marginalisation process<br />

<strong>and</strong> l<strong>and</strong> ab<strong>and</strong>onment are well studied. However little is known about impacts of<br />

environmental factors on ecological succession process <strong>and</strong> related changes in the l<strong>and</strong>scape<br />

spatial structure. Some authors argue that role of these factors is less significant compared to<br />

socio-economic factors (Łowicki 2008; M<strong>and</strong>er et al. 2004).<br />

Though one should bear in mind that once the l<strong>and</strong> has been ab<strong>and</strong>oned, the ecological<br />

succession process takes over <strong>and</strong> environmental factors becomes determinant. The course of<br />

* Corresponding author. Tel.:371 29222324 - Fax: 371 6750 7071<br />

Email address: <strong>and</strong>a.ruskule@bef.lv<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ruskule et al. 2010. Patterns of afforestation process in ab<strong>and</strong>oned agriculture l<strong>and</strong> in Latvia<br />

98<br />

afforestation, its spatial character <strong>and</strong> species composition will depend on factors like soil (its<br />

chemical properties <strong>and</strong> humidity), terrain, initial herb layer, nearby forest, as well as former<br />

l<strong>and</strong> use (Prach et al. 2001a; Prach et al. 2001b; Barth et al 2003; Alard et al. 2005; Daugaviete<br />

2009; Kopecký <strong>and</strong> Vojta 2009; Rosenthal (in press)). Depending on interaction of these factors<br />

ecological succession might be either arrested or stimulated as well as leading to different<br />

succession patterns.<br />

Afforestation of ab<strong>and</strong>oned l<strong>and</strong>scape has significant effect on l<strong>and</strong>scape structure <strong>and</strong> its<br />

ecological functions (Reger et al. 2007; Stoate et al. 2009). Impacts on biodiversity might<br />

varying – at the initial stage of afforestation habitat diversity is increasing, while in long term<br />

perspective l<strong>and</strong>scape becomes more homogenous thus reducing also biological diversity<br />

(Fjellstad et al. 1999; Nikodemus et al. 2005; Sitzia 2010). Ab<strong>and</strong>onment of agricultural l<strong>and</strong> is<br />

reducing areas important for resting, feeding or breeding of migratory birds. However secondary<br />

succession process might be used also for restoration of natural habitats <strong>and</strong> increasing of<br />

abundance of related animal species (Prach et al. 2001a; Stoate et al. 2009).<br />

L<strong>and</strong> ab<strong>and</strong>onment also influence the scenic quality of l<strong>and</strong>scape <strong>and</strong> its identity which might<br />

psychologically deject inhabitants causing feelings like isolation, poverty, shame etc. (Bürgi<br />

2004; Palang et al. 2006; Benjamin 2007). Better underst<strong>and</strong>ing the role on environmental<br />

factors in the process of l<strong>and</strong>scape ecological succession would help to predict the course of its<br />

development <strong>and</strong> to make optimal choice for the future use of the ab<strong>and</strong>oned l<strong>and</strong>.<br />

2. Methodology<br />

The study analyses the spatial character of l<strong>and</strong>scape ecological succession <strong>and</strong> its influencing<br />

environmental factors at the local level within selected pilot areas. The study area comprise an<br />

ab<strong>and</strong>oned agriculture l<strong>and</strong> in Vidzeme, central part of Latvia, which represents typical rural<br />

l<strong>and</strong>scape including scarcely populated areas in the Gauja River valley bordering the strict<br />

reserve zone of the Gauja National Park, as well as more densely populated areas in vicinities of<br />

towns Sigulda, Līgatne <strong>and</strong> Taurene. The pilot areas of the study have been chosen by visual<br />

analysis of the latest available ortophoto images from 2007. The most distinct patterns of<br />

l<strong>and</strong>scape ecological succession within the study area were selected, covering different spatial<br />

character of shrub <strong>and</strong> tree patches <strong>and</strong> their species composition. As result 5 pilot areas have<br />

been chosen representing 4 patterns of l<strong>and</strong>scape ecological succession (see Table 1).<br />

During the field visits actual borders of woody patches have been mapped as well as density <strong>and</strong><br />

composition of woody species have been assessed at the sampling sites of 10x10m, by recording<br />

each tree <strong>and</strong> measuring their height. Sampling sites were selected within each patch, thus their<br />

number was dependent on size of pilot area <strong>and</strong> complexity of the secondary succession pattern<br />

(20 sampling sites in 1st pilot area; 24 - in 2nd pilot area; 26 - in 3rd pilot area; 14 - in 4th pilot<br />

area <strong>and</strong> 15 - in 5th pilot area).<br />

Environmental conditions <strong>and</strong> factors were described in the investigation areas. Altogether 13<br />

soil profiles were described during field works according to the international FAO WRB<br />

classificator (IUSS Working Group WRB, 2007), 57 soil samples were collected from soil<br />

profile diagnostic horizons <strong>and</strong> physical <strong>and</strong> chemical analyses were done in the laboratory<br />

according to the methodology of ICP <strong>Forest</strong> monitoring (Manual on methods…, 2006). Impacts<br />

of nearby forest st<strong>and</strong> <strong>and</strong> its species composition on the character of ecological succession have<br />

been defined by analysing the forest taxation maps obtained from the State <strong>Forest</strong> Service.<br />

Impacts of drainage systems have been analysed using the melioration maps of the former<br />

collective farms. Information on former l<strong>and</strong> use <strong>and</strong> period since agriculture practice has been<br />

given up were obtained from local inhabitants <strong>and</strong> l<strong>and</strong>owners by direct interviews at the pilot<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ruskule et al. 2010. Patterns of afforestation process in ab<strong>and</strong>oned agriculture l<strong>and</strong> in Latvia<br />

99<br />

areas. Dynamics of ecological succession process has been analysed comparing the latest<br />

ortophoto images from 2007 with earlier ones from 1997-1998 <strong>and</strong> 2003-2004.<br />

3. Results<br />

3.1 Development stage <strong>and</strong> character of l<strong>and</strong>scape ecological succession process<br />

Within the study area 4 patterns of l<strong>and</strong>scape ecological succession could be distinguished:<br />

linear development of succession; mosaic development of succession; continuous development<br />

of succession <strong>and</strong> development of succession from edge of a forest (see Figure 1).<br />

1 2<br />

5b<br />

5c<br />

5a<br />

3 4<br />

Figure 1: Patterns of l<strong>and</strong>scape ecological succession: 1; 2 - linear development of succession; 3; 4; 5b -<br />

mosaic development of succession; 5a - continuous development of succession; 5c - development of<br />

succession from edge of a forest.<br />

The most typical pattern in study area is linear succession, where natural afforestation process<br />

starts as linear patches following the ploughing direction. Two representative examples of this<br />

pattern are found in 1st <strong>and</strong> 2nd pilot area. Both areas are placed within larger agriculture fields<br />

(425 ha <strong>and</strong> 376 ha) previously used for crop production. Development of tree cover in these<br />

areas has started ca. 10 years ago (4-5 years after ab<strong>and</strong>onment) <strong>and</strong> it is formed by deciduous<br />

trees - Betula pendula <strong>and</strong> Salix spp. Dominant species within the 1st pilot area is Betula<br />

pendula, forming mostly sparse st<strong>and</strong>s (up to 50 trees per 100m 2 ) with few more dens patches<br />

(up to 126 trees per 100 m 2 ). Height of the trees in these st<strong>and</strong>s reaches 8 m at maximum. The<br />

topography of the area is rather plain with prevailing soil groups: Stagnosols, Arenosols,<br />

Luvisols <strong>and</strong> Phaeozems (according to FAO WRB classification). In 2nd pilot area dominant<br />

are Salix spp. forming visually dense st<strong>and</strong>s, although the number of trees per 100m 2 also mostly<br />

are bellow 50. Height of trees reaches 11 m both for Betula pendula <strong>and</strong> Salix spp. Also the<br />

topography of this area is plain with widely distributed soil groups: Stagnosols <strong>and</strong> Gleysols.<br />

Mosaic l<strong>and</strong>scape succession is represented at 3rd, 4th <strong>and</strong> 5th pilot area. This pattern can be<br />

characterised by higher diversity of woody species <strong>and</strong> irregular shapes of patches. The size of<br />

the surrounding agriculture fields is smaller than in two previous cases (134 ha, 102 ha <strong>and</strong> 38<br />

ha). In 3rd pilot the most common tree species is Betula pendula, however few patches are<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ruskule et al. 2010. Patterns of afforestation process in ab<strong>and</strong>oned agriculture l<strong>and</strong> in Latvia<br />

100<br />

dominated by Picea abies. Tree cover is mostly forming sparse st<strong>and</strong>s. Height of Betula<br />

pendula reaches 10. In 4th pilot area the most widespread species is Picea abies, but also Salix<br />

spp. <strong>and</strong> Betula pendula are frequent. Maximum height of trees is 9 m. In 5th pilot area mosaic<br />

succession pattern is dominated by Betula pendula with rather high share of Picea abies <strong>and</strong><br />

Pinus sylvestris. The highest trees reach 8 m. In this area tree st<strong>and</strong>s are mostly dense (more<br />

than 50 trees per 100 m 2 ). In all these pilot areas tree species cover has started to develop ca. 10<br />

years ago. The topography of the 3rd <strong>and</strong> 4th pilot area is slightly waved, <strong>and</strong> the distribution of<br />

soil cover is rather complex, including Luvisols, Phaeozems, Stagnosols, Gleyosols, Planosols,<br />

as well as Arenosols <strong>and</strong> Podzols. In 5th pilot area the topography is plain <strong>and</strong> mostly covered<br />

by Luvisols, Stagnosols, Gleysols <strong>and</strong> Arenoslos.<br />

Continuous succession pattern has been observed at one part of the 5th pilot area – a narrow<br />

field in earlier years used as arable l<strong>and</strong>, but later regularly mowed. The tree cover has<br />

developed just recently - during the last 2-3 years, forming rather dense st<strong>and</strong> (up to 150 trees<br />

per 100 m 2 ) dominated by Betula pendula up to 3 m height. Development of succession from<br />

the edge of a forest is the most explicit at the 5th pilot area where it is combined with mosaic<br />

succession pattern. However also in 1st <strong>and</strong> 2nd pilot area influence of the forest edge or nearby<br />

forest st<strong>and</strong> on development of linear succession pattern can be observed.<br />

Table 1: Character of l<strong>and</strong>scape ecological succession in pilot areas<br />

Pilot area<br />

1. pilot area village<br />

Nurmiži<br />

2. pilot area near<br />

farmstead “Gobas”<br />

3. pilot area near<br />

village Ieriķi<br />

4. pilot area village<br />

Taurene<br />

5. pilot area near<br />

village Inciems<br />

Pattern of<br />

succession<br />

Year since is<br />

ab<strong>and</strong>oned<br />

Dominant tree species<br />

linear 1996 Betula pendula 65%<br />

from forest edge<br />

Salix spp. 35%<br />

linear 1996 Salix spp. 71%<br />

from forest edge<br />

Betula pendula 29%<br />

mosaic 1995 Betula pendula 74%<br />

Picea abies 18%<br />

Pinus sylvestris 5,4%<br />

mosaic 1995 Picea abies 55%<br />

Salix spp.19%<br />

Betula pendula 14%<br />

continuous; 2000 Betula pendula 62%<br />

mosaic;<br />

Picea abies 23%<br />

from forest edge<br />

Pinus sylvestris 12%<br />

Total area of the<br />

agriculture field (ha)<br />

425<br />

377<br />

135<br />

102<br />

38<br />

3.2 Impact of species composition of surrounding forest on ecological succession process<br />

Unfortunately forest taxation information was not complete (it did not include all forest st<strong>and</strong>s<br />

bordering the pilot areas) therefore comprehensive analysis of relation between species<br />

composition in the forest st<strong>and</strong>s <strong>and</strong> secondary succession patches was not possible. However<br />

from the available information it is obvious that tree species composition not always correspond<br />

between the succession patches <strong>and</strong> the surrounding forest. For example in 3rd pilot area<br />

composition of the dominant species is similar (Betula pendula 50 % - in forest, 74 % - in<br />

succession patches; Picea abies 23 % - in forest, 18 % - in succession patches), while less<br />

dominant species present in forest (e.g. Populus tremula – 13 %; Alnus incana – 11%) can<br />

hardly be found within the succession patches. In the 2nd pilot area where secondary<br />

succession patches are formed by Salix spp. - 71 %, Betula pendula - 29 %, Picea abies - 0,1 %,<br />

Alnus incana - 0,1 %, the species composition of surrounding forest st<strong>and</strong>s is different (Betula<br />

pendula -53 % <strong>and</strong> Populus tremula – 26 %, including also Picea abies - 8%, Alnus incana -<br />

5 % <strong>and</strong> Pinus sylvestris – 3 %).<br />

3.3 Impact of soil conditions on ecological succession process<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ruskule et al. 2010. Patterns of afforestation process in ab<strong>and</strong>oned agriculture l<strong>and</strong> in Latvia<br />

101<br />

Information from soil maps obtained from State L<strong>and</strong> Service as well as from soil sampling<br />

carried out in the pilot areas mainly do not show direct impact of soil groups <strong>and</strong> its chemical<br />

properties on spatial distribution <strong>and</strong> character of the ecological succession patches (see Figure<br />

2). Nevertheless, relationship between soil groups <strong>and</strong> its chemical properties have been<br />

recognized within the 3rd pilot area. There it was established that development of shrubs <strong>and</strong><br />

trees are faster on the soils characterized with poor nutrient status, but on fertile soils this<br />

process could be considerably delayed.<br />

4. Discussion<br />

Figure 2: Soil conditions in pilot area 1 <strong>and</strong> 3.<br />

The study shows that relation between character of l<strong>and</strong>scape ecological succession <strong>and</strong> such<br />

environmental factors as soil properties <strong>and</strong> species composition of surrounding forest are rather<br />

ambiguous, if viewed separately. Development of ecological succession depends on interactions<br />

of various factors like topography, geology, level of ground waters, soil <strong>and</strong> its chemical<br />

properties. The previous l<strong>and</strong> use <strong>and</strong> initial stage of succession prescribed by herbaceous<br />

species colonising the ab<strong>and</strong>oned fields also has significant role in this process (Prach et al<br />

2001b; Kopecký <strong>and</strong> Vojta 2009; Rosenthal (in press)). For example in some cases higher soil<br />

fertility can enhance the growth of shrubs, while in another it might increase density of herb<br />

layer, thus solving down the course of succession by competitive exclusion of trees (Alard et al<br />

2005).<br />

Furthermore distribution of seeds of woody species depends on dominant winds as well as size<br />

<strong>and</strong> configuration of the field. Although some studies highlight importance of distance from the<br />

seed st<strong>and</strong> <strong>and</strong> species composition of the seed st<strong>and</strong> as decisive factor in development of<br />

secondary succession (Daugaviete 2009), we consider that degree of influence of this factor<br />

depends on size of the field as well as heterogeneity of environmental conditions. In very large<br />

<strong>and</strong> fields with strait edges <strong>and</strong> plane topography having uniform environmental conditions the<br />

afforestation process starts form the edge of forest <strong>and</strong> might develop into linear patches, while<br />

in smaller fields with more complex shapes <strong>and</strong> hilly topography, where environmental<br />

conditions are more diverse, the pattern of secondary succession is rather complex, usually<br />

developing as mosaic patches, not always having direct connection to the forest edge. In very<br />

small fields surrounded by forest continuous succession might develop covering simultaneously<br />

all the area in rather short time.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ruskule et al. 2010. Patterns of afforestation process in ab<strong>and</strong>oned agriculture l<strong>and</strong> in Latvia<br />

102<br />

References<br />

Alard D., Chabrerie O., Dutoit T., Roche P., Langlois E., 2005. Patterns of secondary succession<br />

in calcerous grassl<strong>and</strong>s: can we distinguish the influence of former l<strong>and</strong> uses from present<br />

vegetation data Basic <strong>and</strong> Applied Ecology, 6: 161-173.<br />

Benjamin K., Bouchard A., Domon G., 2007. Ab<strong>and</strong>oned farml<strong>and</strong>s as components of rural<br />

l<strong>and</strong>scapes: An analysis of perceptions <strong>and</strong> representations, L<strong>and</strong>scape <strong>and</strong> Urban<br />

Planning, 83: 228-244.<br />

Bartha S., Meiners S.J., Pickett, S.T.A., Cadenasso M.L., 2003. Plant colonization windows in<br />

mesic old field succession, Applied Vegetation Science, 6: 205-212.<br />

Bürgi M., Hersperger A.M, Schneeberger N., 2004. Driving forces of l<strong>and</strong>scape change –<br />

current <strong>and</strong> new directions, L<strong>and</strong>scape Ecology, 19: 857-868.<br />

Daugaviete M., 2009. The qualitative characteristics of Naturally-developed deciduous forest<br />

st<strong>and</strong>s in ab<strong>and</strong>oned agricultural l<strong>and</strong>s. In: Substantiation of deciduous trees cultivation<br />

<strong>and</strong> rational utilisation, new products <strong>and</strong> technologies. State Research Programme, 2005-<br />

2009.Proceedings. Riga, Latvian State Institute of Wood Chemistry: 23-27. (In Latvian)<br />

Fjellstad W.J., Dramstad W.E., 1999. Patterns of change in two contrasting Norwegian<br />

agricultural l<strong>and</strong>scapes, L<strong>and</strong>scape <strong>and</strong> Urban Planning 45: 177 – 191.<br />

IUSS Working Group WRB, 2007. World Reference Base for Soil Resources 2006, first update<br />

2007. World Soil Resources Reports No. 103. FAO, Rome.<br />

Łowicki D., 2008. L<strong>and</strong> use changes in Pol<strong>and</strong> during transformation: Case study of<br />

Wielkopolska region, L<strong>and</strong>scape <strong>and</strong> Urban Planning 87, 279–288.<br />

M<strong>and</strong>er Ü., Palang H., Ihse M., 2004. Development of European l<strong>and</strong>scape, Editorial, L<strong>and</strong>scape<br />

<strong>and</strong> Urban Planning, 67: 1-8.<br />

Manual on methods <strong>and</strong> criteria for harmonized sampling, assessment, monitoring <strong>and</strong> analysis<br />

of the effects of air pollution on forests, (2006) / International Co-operative Programme<br />

on Assessment <strong>and</strong> Monitoring of Air Pollution Effects on <strong>Forest</strong>s / United Nations<br />

Economic Commission for Europe Convention on Long-Range Transboundary Air<br />

Pollution. Part IIIa, Expert Panel on Soil <strong>Forest</strong> Soil Co-ordinating Centre, Research<br />

Institute for Nature <strong>and</strong> <strong>Forest</strong>, Belgium.<br />

Ministry of Agriculture, 2009. <strong>Forest</strong>ry sector in Latvia: 2009.<br />

Nikodemus O., Bell S., Grīne I., Liepiņš I., 2005. The impact of economic, social <strong>and</strong> political<br />

factors on the l<strong>and</strong>scape structure of the Vidzeme Upl<strong>and</strong>s in Latvia, L<strong>and</strong>scape <strong>and</strong><br />

Urban Planning, 70: 57-67.<br />

Palang H., Printsmann A., Konkoly Gyuro E., Urbanc M., Skowronerk E., Woloszyn W., 2006.<br />

The forgotten rural l<strong>and</strong>scapes of Central <strong>and</strong> Eastern Europe, L<strong>and</strong>scape Ecology, 21:<br />

347-357.<br />

Prach K., Bartha S., Joyce C.B., Pyšek P., van Diggelen R., Wiegleb G., 2001 a. The role of<br />

spontaneous vegetation succession in ecosystem restoration: A perspective, Applied<br />

Vegetation Science, 4: 111-114.<br />

Prach K., Pyšek P., van Diggelen R., Bastl M., 2001 b. Spontaneous vegetation succession in<br />

human-disturbed habitats: A pattern across seres, Applied Vegetation Science, 4: 83-88.<br />

Reger B., Otte A., Waldhardt R., 2007. Identifying patterns of l<strong>and</strong>-cover change <strong>and</strong> their<br />

physical attributes in a marginal European l<strong>and</strong>scape, L<strong>and</strong>scape <strong>and</strong> Urban Planning, 81:<br />

104–113.<br />

Rosenthal G., (in press). Secondary succession in a fallow central European wet grassl<strong>and</strong>, Flora.<br />

Sitzia T., Semenzato P., Trentanovi G., 2010 (in press). Natural reforestation is changing spatial<br />

patterns of rural mountain <strong>and</strong> hill l<strong>and</strong>scape: A global overview, <strong>Forest</strong> Ecology <strong>and</strong><br />

Management.<br />

Stoate C., Báldi A., Beja P., Boatman N.D., Herzon I., van Doorn A., de Snoo G.R., Rakosy L.,<br />

Ramwell C., 2009. Ecological impacts of early 21st century agricultural change in Europe<br />

– A review, Journal of Environmental Management, 91: 22-46.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Sayad 2010. Nitrogen retranclocation in pure <strong>and</strong> mixed plantations of Populus deltoides <strong>and</strong> Alnus subcordata<br />

103<br />

Nitrogen retranclocation in pure <strong>and</strong> mixed plantations of Populus<br />

deltoides <strong>and</strong> Alnus subcordata<br />

Ehsan Sayad *<br />

Natural Resources Faculty, Higher Education Complex of Behbahan, Iran<br />

Abstract<br />

Nutrients allocated to green leaves are recycled through 4 parallel pathways: herbivory,<br />

throughfall, foliar resorption <strong>and</strong> litter decomposition. Nutrients recycled trough each pathway<br />

may be of similar magnitude. Populus deltoides <strong>and</strong> Alnus subcordata were planted in five<br />

proportions (100P, 67P:33A, 50P:50A, 33P:67A, 100A) in Noor, Iran. After 7 years, the effects<br />

of species interactions on nutrient retranslocation were assessed. Leaves were collected from the<br />

bottom one-third of the tree. Six representative trees (two near the center of sub-plot <strong>and</strong> one in<br />

each corner of it) of each species were sampled for fully exp<strong>and</strong>ed leaves. Senescent leaves<br />

were also collected from each species in each sub-plot. The Nitrogen retranslocations of both<br />

species were not significantly differed between the different proportions also it was not different<br />

between the two species. Finally, it should be implied that nutrient retranslocation was not<br />

differed as a result of mixing these species at this age.<br />

Keywords: Mixed plantations, Nutrient Retranslocation, Nitrogen fixing tree<br />

1. Introduction<br />

Tree plantations with different purposes are widespread activity all over the world. Almost all<br />

the industrial plantations are monocultures, <strong>and</strong> questions are being raised about the<br />

sustainability of their growth <strong>and</strong> their effects on the site (Khanna 1997). Poplars (Populus L.<br />

spp.) are preferred plantation species, because their fast growth is expected to meet the<br />

extensive dem<strong>and</strong>s of wood for poles, pulp <strong>and</strong> fuel (Kiadaliri 2003; Ghasemi 2000; Ziabari<br />

1993). Repeated harvesting of fast-growing trees such as poplar plantations on short rotations<br />

may deplete site nutrients.<br />

Nitrogen losses are likely to be very important for future growth. Therefore, it is appropriate to<br />

explore new systems of plantation management in which N may be added via fixation (Khanna<br />

1997). Mixed plantation systems seem to be the most appropriate for providing a broader range<br />

of options, such as production, protection, biodiversity conservation, <strong>and</strong> restoration<br />

(Montagnini et al. 1995; Keenan et al. 1995; Guariguata et al. 1995; Parrotta <strong>and</strong> Knowles<br />

1999).<br />

Nutrients allocated to green leaves are recycled through 4 parallel pathways: herbivory (feces<br />

<strong>and</strong> dead bodies), throughfall, foliar resorption <strong>and</strong> litter decomposition. Research often focuses<br />

exclusively on decomposition, but the fraction of nutrients recycled trough each pathway may<br />

be of similar magnitude. This fraction varies with the nutrient considered, the species <strong>and</strong> the<br />

climatic conditions. For example, leaf nitrogen recycling is estimated to be 10% through<br />

herbivory, 5% through throughfall, 40% through resorption <strong>and</strong> 45% through litterfall <strong>and</strong><br />

subsequent decomposition. Locally, the chemical <strong>and</strong> physiological characteristics of leaves<br />

* Corresponding author. Tel.: +989161425493 - Fax: +98(671)2231662<br />

Email address: ehsansaiad@yahoo.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Sayad 2010. Nitrogen retranclocation in pure <strong>and</strong> mixed plantations of Populus deltoides <strong>and</strong> Alnus subcordata<br />

104<br />

may relate to the ecological status of the species (pioneer vs. late successional species, sun vs.<br />

shade tolerant species). Litter chemistry is determined both by the chemistry of green leaves <strong>and</strong><br />

by the nutrient resorption. The latter has not been given enough attention despite its major role<br />

in nutrient conservation both at the individual <strong>and</strong> ecosystem level (Binkley <strong>and</strong> Menyailo<br />

2005).<br />

Thus in this paper, we focus on N resorption in pure <strong>and</strong> mixed plantations in order to increase<br />

our underst<strong>and</strong>ing of the ecology of these tree species in this region. Therefore our question is:<br />

Are there differences in the degree of internal cycling of N (resorption) of each species in<br />

different proportions And dose it differ between the two species<br />

2. Methodology<br />

2.1 Study site<br />

The study area is located at the Chamestan experiment station, in Maz<strong>and</strong>aran province, on the<br />

northern parts of Iran (36º29’N, 51º59’E). Experimental plantations were established in 1996<br />

using a r<strong>and</strong>omized complete block design that included four replicate 40 m × 40 m plots of<br />

each of the following treatments: (i) Populus deltoides (100P), (ii) Alnus subcordata (100A),<br />

(iii) 50% P. deltoides + 50% A. subcordata (50P:50A), (iv) 67% P. deltoides + 33% A.<br />

subcordata (67P:33A), (v) 33% P. deltoides + 67% A. subcordata (33P:67A). Tree spacing<br />

within plantations was 4 m × 4 m <strong>and</strong> tow species were systematicly mixed within rows.<br />

2.3 Leaf<br />

Leaf samples were collected from the st<strong>and</strong>s in September 2003. Leaves were collected from<br />

the bottom one-third of the tree by clipping two small twigs located on opposite sides of the<br />

crown. Six representative trees (two near the center of sub-plot <strong>and</strong> one in each corner of it) of<br />

each species were sampled for fully exp<strong>and</strong>ed leaves. In addition, senescent leaves were<br />

collected from each species in each sub-plot. The samples were dried at 65 o C <strong>and</strong> ground prior<br />

to analysis. Nitrogen was analyzed using the Kjeldhal.<br />

2.5 Statistical Analyses<br />

The percent of retranslocation efficiency was calculated from the fallowing formula (1):<br />

R =<br />

A − B<br />

× 100<br />

A<br />

(1)<br />

In this formula A is the weight of each nutrient in fully exp<strong>and</strong>ed leaf <strong>and</strong> B is the weight of that<br />

nutrient in leaf litter. The percent of retranslocation between the species <strong>and</strong> for each species in<br />

different treatments were compared using general linear model analysis of variance (ANOVA)<br />

tests. Normality of the data was checked for analyses. Statistical analyses were done using SAS<br />

9 software.<br />

3. Result<br />

The Nitrogen retranslocations of both species were not significantly differed between the<br />

different proportions. Alnus Nitrogen retranslocation was higher in pure plantations than the<br />

mixtures, but for Populus it was higher under 67%P: 33%A than the other treatments. Nitrogen<br />

retranslocation also was not different between the two species, whereas the retranslocation in<br />

Alnus as Nitrogen fixing tree was lower than Populus (Figure 1).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Sayad 2010. Nitrogen retranclocation in pure <strong>and</strong> mixed plantations of Populus deltoides <strong>and</strong> Alnus subcordata<br />

105<br />

Figure 1: N retranslocation of each species in mixed plantations <strong>and</strong> between the two species in pure<br />

plantations<br />

4. Discussion<br />

In line with our results, Parrotta (1999) showed that N retranslocation were not significantly<br />

different among treatments for Eucalyptus. He implied that this result suggest that the<br />

association with either Nitrogen fixing species is not (yet) having any positive influence on<br />

nutrient availability of Eucalyptus. The same conclusion could be stated for the current results.<br />

It means that the species have not any effect on nitrogen cycling of other one. In contrary to our<br />

results about the two species Cuevas <strong>and</strong> Lugo (1998) found different nitrogen retrenslocation<br />

between the species. In this case it could be concluded that the nitrogen requirements of these<br />

two species at this age are similar, although Alnus had the lower retranclocation. Finally, it<br />

should be implied that nutrient retranslocation was not differed as a result of mixing these<br />

species at this age.<br />

References<br />

Binkley, D., Menyailo, O., 2005. Tree Species Effects on Soil: Implications for <strong>Global</strong> <strong>Change</strong>,<br />

NATO Science Series. IV Earth <strong>and</strong> Environmental Sciences-Vol.55. 358 pp.<br />

Cuevas, E., <strong>and</strong> Lugo, A.E., 1998. Dynamics of organic matter <strong>and</strong> nutrient return from litterfall<br />

in st<strong>and</strong>s of ten tropival tree plantation species. <strong>Forest</strong> Ecology <strong>and</strong> Management,<br />

112:263-279.<br />

Ghasemi, R., 2000. Study of phenology of different Populus in Karaj <strong>and</strong> Safrabasteh Gilan.<br />

M.Sc. thesis of Tarbiat Modarres University. 171 p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Sayad 2010. Nitrogen retranclocation in pure <strong>and</strong> mixed plantations of Populus deltoides <strong>and</strong> Alnus subcordata<br />

106<br />

Guariguata, M.R., Rheingans, R. <strong>and</strong> Montagnini, F., 1995. Early woody invasion under tree<br />

plantations in Costa Rica: implications for forest restoration. Restoration Ecology, 3(4):<br />

252–260.<br />

Keenan, R.J., Lamb, D. <strong>and</strong> Sexton, G., 1995. Experience with mixed species rainforest<br />

plantations in North Queensl<strong>and</strong>. Commonwealth <strong>Forest</strong>ry Review, 74(4): 315–321.<br />

Khanna, P.K., 1997. Comparison of growth <strong>and</strong> nutrition of young monocultures <strong>and</strong> mixed<br />

st<strong>and</strong>s of Eucalyptus globules <strong>and</strong> Acacia mearnsii. <strong>Forest</strong> Ecology <strong>and</strong> Management,<br />

94(1–3): 105–113.<br />

Kiadaliri, Sh., 2003. Study of populus plantations on different soils in western parts of<br />

Maz<strong>and</strong>aran. M.Sc. thesis of Tarbiat Modarres University. 94 p.<br />

Montagnini, F., Gonzalez, E., Rheingans, R. <strong>and</strong> Porras, C., 1995. Mixed <strong>and</strong> pure forest<br />

plantations in the humid neotropics: a comparison of early growth, pest damage <strong>and</strong><br />

establishment costs. Commonwealth <strong>Forest</strong>ry Review, 74(4): 306–314.<br />

Parrotta, J.A., 1999. Productivity, nutrient cycling, <strong>and</strong> succession in single- <strong>and</strong> mixed-species<br />

plantations of Casuarina equisetifolia, Eucalyptus robusta, <strong>and</strong> Leucaena leucocephala in<br />

Puerto Rico. <strong>Forest</strong> Ecology <strong>and</strong> Management, 124(1): 45–77.<br />

Parrotta, J.A. <strong>and</strong> Knowles, O.H., 1999. Restoration of tropical moist forests on bauxite-mined<br />

l<strong>and</strong>s in the Brazilian Amazon. Restoration Ecology, 7(2): 103–116.<br />

Ziabari, Z.F., 1993. The importance of Populus in forestry. Proceedings of forest restoration<br />

<strong>and</strong> forestry. Research Institute of <strong>Forest</strong>s <strong>and</strong> Range L<strong>and</strong>s publication, Iran. 11 p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Zhelezov 2010. Evaluation of the ecosystem service in the forest formations of Biosphere Reserve “Srebarna”<br />

107<br />

Evaluation of the ecosystem service in the forest formations of<br />

Biosphere Reserve “Srebarna”, Northeastern Bulgaria<br />

Georgi Zhelezov *<br />

Institute of Geography, Bulgarian Academy of Sciences, Bulgaria<br />

Abstract<br />

The investigation observes the basic parameters in ecosystem services of the forest formation on<br />

the territory of Biosphere Reserve „Lake Sreburna”. Wetl<strong>and</strong> system is a part of Danube<br />

catchments. Importance of the lake depends of rich biological <strong>and</strong> l<strong>and</strong>scape diversity. The most<br />

important are bird colonies of Dalmatian pelican <strong>and</strong> waterfowl birds.<br />

<strong>Forest</strong> formations in the lake <strong>and</strong> around the wetl<strong>and</strong> systems are natural place for nesting <strong>and</strong><br />

spreading of the bird populations. They are dominated of willows <strong>and</strong> popular species. The<br />

forest formations in Danubian isl<strong>and</strong>s have high level of importance as quality <strong>and</strong> diversity of<br />

the ecosystem services. They are flooded for the different period of time during the high Danube<br />

waters. This process reflects <strong>and</strong> determines the complex of ecosystem services in the reserve.<br />

The results of the research can use for optimization of nature protection <strong>and</strong> conservation<br />

activities in the Biosphere Reserve „Srebarna” <strong>and</strong> the territories around the wetl<strong>and</strong> system.<br />

Keywords: ecosystem services, forest formation, evaluation<br />

Introduction<br />

Ecosystems provide different ecosystem services. The concept of ecosystem services deals with<br />

the benefits that humans gain from natural processes <strong>and</strong> ecological functions (Costanza et al.<br />

1997). Ecosystem services are supporting (nutrient cycling, soil formation, primary production<br />

etc.), provisioning – referring to the resource supply (food, fuel, fiber, water, other goods),<br />

regulatory (associated with climate, atmosphere, extreme events, water quality etc.) <strong>and</strong> cultural<br />

(aestetic, spiritual, educational, recreational etc.). Publication of Millenium Ecosystem<br />

Assessment in 2005 focused on ecosystem services of different ecosystems, stressing their<br />

overall decline.<br />

<strong>Forest</strong> formations are among the most endangered systems in spite of the fact that they have a<br />

great influence on human wellbeing <strong>and</strong> are providing to us the majority of important<br />

ecosystem services. <strong>Forest</strong> formations of Danubian isl<strong>and</strong> <strong>and</strong> around the lake are marked by<br />

changes in water level during the year. Water level fluctuations create a variety of habitats with<br />

diverse communities. Habitats are delineated by a range of changes in water regime, soil<br />

properties <strong>and</strong> some other factors. Many primary producers are well adapted to these changes<br />

(avoiding unfavourable conditions, being cosmopolites, having persistent stadia or amphibious<br />

character) <strong>and</strong> find optimal conditions for their survival while supporting a variety of other<br />

organisms. Water level during the vegetation period as well as the intensity, timing <strong>and</strong> the<br />

extent of floods influence primary production <strong>and</strong> other processes i.e. mineralization,<br />

decomposition, colonization with plants, as revealed from different studies. The researches<br />

revealed that ecosystem services depend on biotic community complexity, especially on<br />

vegetation type of the area.<br />

* Corresponding author. Tel.: +359 2 979 6309 - Fax: +359 2 870 0204<br />

Email address: gzhelezov@abv.bg; zhelezov75@yahoo.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Zhelezov 2010. Evaluation of the ecosystem service in the forest formations of Biosphere Reserve “Srebarna”<br />

108<br />

The problems of evaluation of the ecosystem services in the wetl<strong>and</strong> nature system are<br />

presented in the investigations of Br<strong>and</strong>er et al. (2006), Hoehn J et al. (2003), Heong K.L.<br />

(2008) etc.<br />

Object<br />

The lake Srebarna is one of the biggest wetl<strong>and</strong>s along the Bulgarian sector of river Danuve<br />

(Table 1). It is situated in northeastern Bulgaria, near the village of the same name, 18 km west<br />

of Silistra <strong>and</strong> 2 km south of the Danube. The reserve embraces 6 km² of protected area <strong>and</strong> a<br />

buffer zone of 5.4 km². Diversity of the reserve <strong>and</strong> surrounding territories is determined in<br />

Corine Biotope Project in 13 habitate types. The importance of the wetl<strong>and</strong> system is determine<br />

from diversity of the rare <strong>and</strong> protected bird species as Dalmatian Pelican (Pelicanus crispus),<br />

Mute Swan (Cygnus olor), six heron species (Ardea alba, Ardea cinerea, Ardea purpurea,<br />

Ardeola raloides, Egretta garzetta etc), cormorants (Phalacrocorax carbo <strong>and</strong> Halietor<br />

pygmaeus) etc.<br />

Table 1: Limnological characteristics (Management plan, 2001)<br />

Indicator<br />

Value<br />

Elevation (m) 10,0 - 13,2<br />

Water catchments (km 2 ) 402<br />

Length of shore line (km) 18,5<br />

Reserve area (ha) 902,1<br />

Open water body (ha) 120<br />

Capacity (km 3 ) – low water levels 2,81<br />

Capacity (km 3 ) – high water levels 14,35<br />

Maximal depth (m) 3,3<br />

Years flow (m 3 ) 12,48<br />

Time of stay (months) 2,67<br />

Diversity of biogeographical provinces is determined from three provinces based on Udvardy<br />

(1975):<br />

- Middle European forest formations;<br />

- Pontiac steps formations;<br />

- Mountain territories.<br />

Ecosystem diversity in the reserve “Srebarna” is determined from different wetl<strong>and</strong> types based<br />

on the Ramsar convention (1971):<br />

M Permanente rivers – river Danube between right bank <strong>and</strong> isl<strong>and</strong> Devnia;<br />

O Permanente fresh water lake – open water body;<br />

R Seasonal marshes <strong>and</strong> water bodies – area between left bank of river Danube<br />

<strong>and</strong> river dike from Silistra to village Vetren;<br />

X f Fresh water bodies with domination of forest formations; seasonal flooded<br />

forests – whole territory of isl<strong>and</strong> Devnia <strong>and</strong> part of the river bank between Silistra <strong>and</strong> village<br />

Vetren <strong>and</strong> right bank of river Danube.<br />

X k Underground karst <strong>and</strong> cave hydrological systems – natural spring<br />

“Kanarichkata” in the south part of the reserve.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Zhelezov 2010. Evaluation of the ecosystem service in the forest formations of Biosphere Reserve “Srebarna”<br />

109<br />

Bialata prust<br />

∗<br />

Devnia<br />

Danube<br />

Kodzha Bair<br />

(84,1 m)<br />

Kara Burun<br />

(111 m)<br />

Srebarna<br />

- w ater body<br />

- hydrophites formation<br />

- forest formation w ith<br />

domination of Pinus nigra<br />

- forest formation w ith<br />

domination of Robinia pseudoacacia<br />

- forest formation w ith<br />

domination of Quercus species<br />

- forest formation w ith<br />

domination of Populus species<br />

- f orest f ormation with<br />

domination of Salix alba<br />

<strong>and</strong> Populus alba<br />

- grass formation<br />

- flooded area<br />

- agriculture area<br />

- villages<br />

Figure 1: <strong>Forest</strong> formations in the Biosphere Reserve “Srebarna”<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Zhelezov 2010. Evaluation of the ecosystem service in the forest formations of Biosphere Reserve “Srebarna”<br />

110<br />

Research problem<br />

The topic of the research is to investigate the importance of forest formations on the territory of<br />

Biosphere reserve Srebarna (Northeastern Bulgaria) from the point of view of ecosystem<br />

services. Basic vegetation types (Fig. 1), which are presented on the territory of the reserve are:<br />

1. Hydrophite <strong>and</strong> hygrophite formations dominated from Phragmites australis, Typha<br />

angustifolia <strong>and</strong> Typha latipholia, Schoenoplectus lacustris, Sch. triquetra, Sch.<br />

Tabernemontana, Salix cinerea etc.<br />

2. Mixed forest formations dominated from Populus alba <strong>and</strong> Salix alba.<br />

3. Mezokserotermal grass formation dominated from Poa bulbosa, Lolium perenne,<br />

Cynodon dactilon, Dichantium ischaemum, Chrysopogon gryllus.<br />

4. Mixed forest formations dominated from Quercus cerris, Quercus pubescens, Quercus<br />

virgiliana.<br />

5. Mixed forest formations dominated from Quercus frainetto <strong>and</strong> Carpinus orientalis<br />

with Mediterannian elements.<br />

6. <strong>Forest</strong> formation of Tilia argentea mixed with Salix alba.<br />

7. Artificial formations of Robinia pseudoacacia <strong>and</strong> Gleditsia triacantha.<br />

8. Artificial formations of Pinus nigra.<br />

9. Agricultural areas on the place forest formations from Quercus cerris <strong>and</strong> Quercus<br />

virgiliana, in some places mixed with Quercus pedunculiflora.<br />

Results<br />

Ecosystem services of the forest formations are connected with specific of the l<strong>and</strong>scape<br />

diversity <strong>and</strong> determinate sustainability of the ecosystems. They secured food provision <strong>and</strong><br />

nesting condition for the bird species. They have role in the ecological balance of the<br />

hydrological resources <strong>and</strong> whole process of interaction between species. There is real<br />

opportunity for using of the biomass in energy producing. <strong>Forest</strong> formations determine<br />

attractiveness of the territory <strong>and</strong> circumstances for development of tourism.<br />

Present research includes investigation of the key forest formations in Biosphere Reserve<br />

“Srebarna”.<br />

The first research territory included investigation of artificial forest formations of Robinia<br />

pseudoacacia <strong>and</strong> Gleditsia triacantha on the place of the vineyard terraces along the slope of<br />

hill Kodzha bair (84,1 m) in west part of the reserve. The region characterizes with high level of<br />

the erosion processes. The age of forest is 32 years. Whole quantity of the tree phytomass is<br />

75,7 t/ha. The low level of phytomass of forest formation is determined from the fact that these<br />

plant species are not typical for the region. The analysis in the management plan of the reserve<br />

recommends replacing of the artificial forest formation with natural Danubian forest formations<br />

of Tilia argentea mixed with Salix alba.<br />

The second research area is situated on the foot of hill Kodzha Bair (13-14 m) <strong>and</strong> showed<br />

hydrophytes formations dominated Phragmites australis, Typha angustifolia, Typha latipholia,<br />

Schoenoplectus lacustris <strong>and</strong> mixed with Salix cinerea. The productivity of reed formations is<br />

2,9 t/ha. The hydrophytes species have very important transpiration role in hydrological regime<br />

of the wetl<strong>and</strong> system The intensity of transpiration is 1,79 gr/dm 2 /h for Populus alba, 0,575<br />

g/dm 2 /h for Salix cinerea <strong>and</strong> 0,478 gr/dm 2 /h for Phragmites australis. Populus formations in<br />

the buffer zone (18, 1 ha) transpirated 70 200 t water for one season. Populus formations in the<br />

protected area (56 ha) transpirated 218 400 t water for one season. Hydrophytes formations in<br />

the reserve (402 ha) transpirated 1 413 000 t water for one season (Management plan, 2001).<br />

The third research area is situated in the west part of isl<strong>and</strong> Devnia (10-11 m), which is<br />

part of the reserve territory. <strong>Forest</strong> formation is dominated of Populus alba in the central<br />

part of the isl<strong>and</strong> <strong>and</strong> Salix alba in the periphery, mixed with Gleditsia triacantha. The<br />

phytomass of Salix alba is 215 t/ha, Gleditsia triacantha - 0,97 t/ha <strong>and</strong> Populus alba - 6,3<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Zhelezov 2010. Evaluation of the ecosystem service in the forest formations of Biosphere Reserve “Srebarna”<br />

111<br />

t/ha in the periphery, where density of this species is not high. The isl<strong>and</strong> territory saved<br />

<strong>and</strong> protected the classical Danubian formation with low anthropogenic impact <strong>and</strong> can<br />

be use as etalon territory.<br />

The fourth research area is situated in east part of the reserve along the slope of hill<br />

Kara Burun (111 m). The forest formations are presented with mixed forest formations<br />

diminated from Quercus cerris, Quercus pubescens, Quercus virgiliana <strong>and</strong> mixed forest<br />

formations diminated from Quercus frainetto <strong>and</strong> Carpinus orientalis. The low quantity of tree<br />

phytomass (67,8 t/ha) is depended from low density of the formations <strong>and</strong> high<br />

anthropogenic impact.<br />

The ecosystem service of forest formations in the Biosphere Reserve „Srebarna” can<br />

observe in several aspects:<br />

- supporting - primary production, nesting conditions;<br />

- provisioning – food, wood, <strong>and</strong> other good;<br />

- regulatory – water quality <strong>and</strong> water regime, regulation of erosion processes;<br />

- cultural – aestetic, recreational <strong>and</strong> touristical.<br />

Conclusions<br />

The basic problem in the ecological situation of Srebarna wetl<strong>and</strong> system is intensive<br />

degradation of the systems as a result of long-time anthropogenic impact in the region –<br />

building of river dikes <strong>and</strong> irrigation system.<br />

The eutrophication processes are connected with productivity of the vegetation formations.<br />

Regulation of the productivity of the hydrophite formations in the lake is basic aim of the<br />

present nature protected activities <strong>and</strong> management plans of the wetl<strong>and</strong> systems.<br />

<strong>Forest</strong> formations from the point of view of ecosystem services have important role in the<br />

protection of the wetl<strong>and</strong> system from slope erosion in south, east <strong>and</strong> especially west part of<br />

the wetl<strong>and</strong> system. Some forest formation (Salix cinerea) combined with hydrophite<br />

formations are key factor for degradation of the lake (closing of water body, increasing of<br />

bottom substrate etc). Mixed forest formation of Salix alba <strong>and</strong> Populus alba are important<br />

habitat for the nesting colonies of waterfowl birds, especially cormorant <strong>and</strong> heron species.<br />

Development of monitoring system for hydrophite formations <strong>and</strong> water regime is basic<br />

problem connected with future investigations <strong>and</strong> management of the wetl<strong>and</strong>s system.<br />

Acknowledgement<br />

This work is supported by the European Social Fund <strong>and</strong> Bulgarian Ministry of Education,<br />

Youth <strong>and</strong> Science under Operative Program “Human Resources Development”, Grant<br />

BG051PO001-3.3.04/40.<br />

References<br />

Br<strong>and</strong>er L., Raymond J.G., Florax M, Vermaat J., 2006. Empirics of Wetl<strong>and</strong> Valuation:<br />

Comprehensive Summary <strong>and</strong> a Meta-Analysis of the Literature. Environmental &<br />

Resource Economics 33: 223–250.<br />

Costanza et al., 1997 The value of the world's ecosystem services <strong>and</strong> natural capital. Nature<br />

387:253-260.<br />

Management plan of reserve Srebarna, 2001.<br />

Heong K.L., 2008 Biodiversity, ecosystem services <strong>and</strong> pest management. Second International<br />

Plantational Industry Conference <strong>and</strong> Exibition, Shah Alam.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Zhelezov 2010. Evaluation of the ecosystem service in the forest formations of Biosphere Reserve “Srebarna”<br />

112<br />

Hoehn J., Lupi F., Kaplowitz M., 2003 Untying a Lancastrian bundle: valuing ecosystems <strong>and</strong><br />

ecosystem services for wetl<strong>and</strong> mitigation. Journal of Environmental Management 68<br />

263–272.<br />

Ramsar convention, 1971.<br />

Udvardy, M. D. F., 1975. A classification of the biogeographical provinces of the world. IUCN<br />

Occasional Paper no. 18. Morges, Switzerl<strong>and</strong>: IUCN.<br />

Udvardy, Miklos D. F., 1975 "World Biogeographical Provinces" (Map). The CoEvolution<br />

Quarterly, Sausalito, California.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Section 3<br />

Disturbances in changing l<strong>and</strong>scapes


A. Altamirano et al. 2010. Human-caused forest fire in Mediterranean ecosystems of Chile<br />

114<br />

Human-caused forest fire in Mediterranean ecosystems of Chile:<br />

modelling l<strong>and</strong>scape spatial patterns on forest fire occurrence<br />

Adison Altamirano 1* , Christian Salas 1,2 & Valeska Yaitul 2<br />

1 Departamento de Ciencias <strong>Forest</strong>ales, Universidad de La Frontera, Temuco, Chile<br />

2 School of <strong>Forest</strong>ry <strong>and</strong> Environmental Studies, Yale University, USA<br />

Abstract<br />

Modelling forest fire occurrence has been lacked for the southern hemisphere, in particular for<br />

Chilean ecosystems. We developed models to investigate the relationship between forest fire<br />

occurrence <strong>and</strong> l<strong>and</strong>scape heterogeneity in Mediterranean ecosystems of Chile. We used<br />

georreferenced forest fires data from 2004 to 2008. Our data of spatial heterogeneity were<br />

obtained at multiple spatial scales, including climatic, topographic, human-related, <strong>and</strong> l<strong>and</strong>cover<br />

variables. We fit a logistic model in order to predict forest fire occurrence as a function of<br />

our potential predictor variables. Our best model suggests that the probability of forest fires<br />

occurrence is related to high both temperature <strong>and</strong> precipitation, <strong>and</strong> lower distance to cities.<br />

Our predictions suggest that 46% of the study area has high probability of forest fires<br />

occurrence. Our findings can help to take adequate decisions regarding l<strong>and</strong> planning.<br />

Keywords: Disturbance, logistic regression, remote sensing, l<strong>and</strong> planning.<br />

1. Introduction<br />

Fire disturbance is recognized as an important problem because it can devastate natural<br />

resources, human property, <strong>and</strong> threaten human lives (Cardille et al. 2001; Gustafson et al.<br />

2004; González et al. 2005; Ryu et al. 2007). <strong>Forest</strong> fires result in enormous economic losses<br />

because they affect environmental, recreational, <strong>and</strong> amenity values as well as consume timber,<br />

degrade real estate, <strong>and</strong> generate high cost of suppression (Yadav <strong>and</strong> Kaushik 2007).<br />

Modelling forest fire occurrence (i.e., where <strong>and</strong> when a forest fire starts) has recently been<br />

conducted in the northern hemisphere (Calef et al. 2008; Lozano et al. 2007; Ryu et al. 2007;<br />

Vega-Garcia <strong>and</strong> Chuvieco 2006), however, efforts on the subject are lacking for the southern<br />

hemisphere, in particular for Chilean ecosystems. Some studies in Chile have focused in postfire<br />

effects on vegetation dynamics (Navarro et al. 2008; Litton <strong>and</strong> Santelices 2003), but<br />

studies on predicting forest fires occurrence are lacking. <strong>Forest</strong> fire occurrence has increased in<br />

the last years in Chile, being the mean frequency about five thous<strong>and</strong> forest fires per year. These<br />

fires have affected a mean area of about 500 km2 per year (Navarro et al. 2008; CONAF 2009)<br />

being the human activity the main cause of fire ignition (CONAF 2009). We developed models<br />

to investigate the relationship between forest fire occurrence <strong>and</strong> l<strong>and</strong>scape heterogeneity in<br />

Mediterranean ecosystems of Chile.<br />

* Corresponding author. Tel.: +56 45 325658 - Fax: +56 45 325634<br />

Email address: aaltamiranofro.cl<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Altamirano et al. 2010. Human-caused forest fire in Mediterranean ecosystems of Chile<br />

115<br />

2. Methodology<br />

The study area extends over 892 km 2 <strong>and</strong> is located in Eastern Central Chile covering parts of<br />

the Valparaíso <strong>and</strong> Metropolitan regions (See Figure 2 a). We selected a l<strong>and</strong>scape with<br />

temporal-stability in composition <strong>and</strong> no other significant processes than fire <strong>and</strong> succession<br />

operating at the l<strong>and</strong>scape level (Vega-García <strong>and</strong> Chuvieco 2006).<br />

We used georreferenced forest fires data from a 5-year period of fire occurrence from 2004 to<br />

2008. The database consisted of 7,210 observations, out of which 891 were pixels that burned<br />

between 2004 <strong>and</strong> 2008. A distance of 25 x 25 pixels (750 m) was used to compute the cooccurrence<br />

matrices, since small windows result in very sparse matrices.<br />

Our data of spatial heterogeneity were obtained at multiple spatial scales, including climatic,<br />

topographic, human-related, <strong>and</strong> l<strong>and</strong>-cover variables from satellite imagery.<br />

We fit a logistic model in order to predict forest fire occurrence as a function of our potential<br />

predictor variables. The relationship modeled was that between the binary response variable<br />

(one = burned, zero = not burned) <strong>and</strong> the predictor variables.<br />

3. Result<br />

We selected the best’s no correlated predictor variables. Temperature correlated strongly with<br />

Elevation (r = 0.75) (See Figure 1 a), so only one of the two variables could be used in the same<br />

model. So, we selected Temperature in order to be more biologically meaningful. No significant<br />

correlation was found between Temperature <strong>and</strong> Precipitation (r = 0.31) (See Figure 1 b). Our<br />

best model suggests that the probability of forest fires occurrence is related to high both<br />

temperature <strong>and</strong> precipitation, <strong>and</strong> lower distance to cities (Table 1). The high probability of<br />

forest fire occurrence related to high precipitation could be unusual. However, it can be<br />

explained because the most precipitation is concentrated in the first 500 masl because the local<br />

topographic conditions (See Figure 1 c). Our predictions suggest that 46% (410 km2) of the<br />

study area has high probability of forest fires occurrence, being concentrated in the eastern<br />

locations of the study area (See Figure 2 b). Our model correctly classified about 73% of our<br />

validation dataset.<br />

4. Discussion<br />

The information of this study may be useful for hazard reduction, indicating that risk of forest<br />

fire occurrence (Ryu et al. 2007; Vega-Garcia <strong>and</strong> Chuvieco 2006). Study area is one of the<br />

most populated regions of Chile. Therefore, our findings can help to take adequate decisions<br />

regarding to l<strong>and</strong> <strong>and</strong> urban planning.<br />

If climate determines patterns of forest fire occurrence, then when the climatic variables change,<br />

forest fire occurrence should change. This might have important consequences for long-term<br />

l<strong>and</strong> <strong>and</strong> urban planning, since prioritization of high probability of forest fire occurrence today<br />

might not be effective for the future in the face of climate change.<br />

Exploring new statistical model approach would allow to improving the predictive capability of<br />

the models. So, part of our future research will target to this subject.<br />

References<br />

Calef, M.P., McGuire, A.D., <strong>and</strong> Chapin, F.S., 2008. Human Influences on Wildfire in Alaska<br />

from 1988 through 2005: An Analysis of the Spatial Patterns of Human Impacts. Earth<br />

Interactions, 12 (1): 1–17.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Altamirano et al. 2010. Human-caused forest fire in Mediterranean ecosystems of Chile<br />

116<br />

Cardille, J.A., Ventura, S.J., <strong>and</strong> Turner, M.G., 2001. Environmental <strong>and</strong> social factors<br />

influencing wildfires in the Upper Midwest, United States. Ecological Applications, 11:<br />

111–127.<br />

CONAF., 2009. Corporación Nacional <strong>Forest</strong>al. Recursos <strong>Forest</strong>ales. Protección contra<br />

incendios forestales. http://www.conaf.cl, last accessed 9 june 2008.<br />

González, M., Veblen, T.T., <strong>and</strong> Sibold, J.S., 2005. Fire history of Araucaria–Nothofagus<br />

forests in Villarrica National Park, Chile. Journal of Biogeography, 32: 1187–1202.<br />

Gustafson, E.J., Zollner, P.A., Sturtevant, B.R., He, H.S., <strong>and</strong> Mladenoff, D.J., 2004. Influence<br />

of forest management alternatives <strong>and</strong> l<strong>and</strong> type on susceptibility to fire in northern<br />

Wisconsin, USA. L<strong>and</strong>scape Ecology, 19: 327–341.<br />

Litton, C.M. <strong>and</strong> Santelices, R., 2003. Effect of wildfire on soil physical <strong>and</strong> chemical<br />

properties in a Nothofagus glauca forest, Chile. Revista Chilena de Historia Natural, 76:<br />

529-542.<br />

Lozano, F.J., Suárez-Seoane, S., <strong>and</strong> de Luis, E., 2007. Assessment of several spectral indices<br />

derived from multi-temporal L<strong>and</strong>sat data for fire occurrence probability modelling.<br />

Remote Sensing of Environment, 107: 533–544.<br />

Navarro, R.M., Hayas, A., García-Ferrer, A., Hernández, R., Duhalde, P., <strong>and</strong> González, L.,<br />

2008. Caracterización de la situación posincendio en el área afectada por el incendio de<br />

2005 en el Parque Nacional de Torres del Paine (Chile) a partir de imágenes<br />

multiespectrales. Revista Chilena de Historia Natural, 81: 95-110.<br />

Ryu, S., Chen, J., Zheng, D. <strong>and</strong> Lacroix, J.J., 2007. Relating surface fire spread to l<strong>and</strong>scape<br />

structure: An application of FARSITE in a managed forest l<strong>and</strong>scape. L<strong>and</strong>scape <strong>and</strong><br />

Urban Planning, 83: 275–283.<br />

Vega-García, C. <strong>and</strong> Chuvieco, E., 2006. Applying local measures of spatial heterogeneity to<br />

L<strong>and</strong>sat-TM images for predicting wildfire occurrence in Mediterranean l<strong>and</strong>scapes.<br />

L<strong>and</strong>scape Ecology, 21:595–605.<br />

Yadav, B. <strong>and</strong> Kaushik, R., 2007. <strong>Forest</strong> fire <strong>and</strong> its impacts. Plant Archives, 7 (1): 33-37.<br />

Table 1: Best model for predicting spatial variation of forest fire occurrence in the study area.<br />

Variables Coefficients St<strong>and</strong>ard error z value<br />

Intercept -9.2939 1.6625 -5.54*<br />

Annual mean temperature 0.0437 0.0097 4.43*<br />

Distance to cities -0.0001 0.0000 -7.26*<br />

Annual precipitation 0.0090 0.0009 9.60*<br />

* P < 0.0001<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Altamirano et al. 2010. Human-caused forest fire in Mediterranean ecosystems of Chile<br />

117<br />

a)<br />

Temperature (C o )<br />

10 14<br />

0 500 1000 1500 2000<br />

Elevation (masl)<br />

b)<br />

Temperature (C o )<br />

10 14<br />

300 400 500 600 700<br />

Precipitation (mm)<br />

c)<br />

Precipitation (mm)<br />

300 500 700<br />

0 500 1000 1500 2000<br />

Elevation (masl)<br />

Figure 1: Relationship between predictor variables of forest fire occurrence in the study area. a) Annual<br />

mean temperature <strong>and</strong> Elevation; b) Annual mean temperature <strong>and</strong> Annual precipitation; <strong>and</strong> c) Annual<br />

precipitation <strong>and</strong> Elevation.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Altamirano et al. 2010. Human-caused forest fire in Mediterranean ecosystems of Chile<br />

118<br />

Figure 2: a) Map of study area <strong>and</strong> records of forest fire between 2004 <strong>and</strong> 2008, <strong>and</strong> b) Map of forest fire<br />

occurrence probability based on the predictive model.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Curt & J. Pausas 2010. Are changes in fire regime threatening cork oak-shrubl<strong>and</strong> mosaics<br />

119<br />

Are changes in fire regime threatening cork oak-shrubl<strong>and</strong> mosaics<br />

Thomas Curt 1* & Juli Pausas 2<br />

1 Cemagref, UR EMAX Ecosystèmes méditerranéens et Risques,<br />

3275 route Cézanne - CS 40061 - 13182 Aix-en-Provence cedex 5, France<br />

2 CIDE, CSIC Apartado Oficial, Camí de la Marjal s/n 46470 Albal, Valencia, Spain<br />

Abstract<br />

We simulated the changes of vegetation abundance <strong>and</strong> richness among mosaics of cork oak<br />

woodl<strong>and</strong>s <strong>and</strong> shrubl<strong>and</strong>s using the LASS-Fatel<strong>and</strong> model (Pausas <strong>and</strong> Ramos, 2004). During<br />

simulations, similar mosaics have been submitted to different fire regimes including fire<br />

recurrence, fire size, <strong>and</strong> fire severity. The results showed that cork oak populations are stable<br />

under such fire regimes, suggesting that the shift from cork oak to pure shrubl<strong>and</strong>s are likely to<br />

be due to a combination of disturbances by fires <strong>and</strong> by droughts.<br />

Keywords: Shrubl<strong>and</strong>, cork oak (Quercus suber L.), fire regime, LASS Fatel<strong>and</strong>, Mediterranean<br />

l<strong>and</strong>scape<br />

1. Introduction<br />

Mosaics of shrubl<strong>and</strong>s <strong>and</strong> trees are a common feature in the Mediterranean fire-prone<br />

ecosystems such as in cork oak woodl<strong>and</strong>s associated with so-called maquis (Aronson et al.,<br />

2009). In these ecosystems, trees interact strongly with shrubs in the context of disturbance by<br />

recurrent wildfires. Actually, shrubby understory is a key factor determining the interval<br />

between successive fires <strong>and</strong> the intensity of fires as shrubs are flammable fuels that can rebuild<br />

rapidly after disturbance (Baeza et al., 2006). The composition, cover <strong>and</strong> flammability of<br />

shrubby fuels may lead to a high mortality of mature woody seeders (Moreira et al., 2007), but<br />

also limit drastically the resprouting of surviving stems (Pausas, 1997) <strong>and</strong> the establishment of<br />

young individuals from seeds (Curt et al., 2009).<br />

Cork oak (Quercus suber) is renowned as especially fire-resistant <strong>and</strong> fire-resilient (Pausas,<br />

1997). However, the legally-protected cork oak ecosystems experience increasing tree mortality<br />

<strong>and</strong> regeneration failure in the Maures massif (southern France), likely due to recurrent wildfires<br />

<strong>and</strong> severe summer droughts. This is hypothesized to cause major population impacts in a<br />

context of climate change, as fire recurrence <strong>and</strong> fire severity should increase. L<strong>and</strong> <strong>and</strong> forest<br />

managers need the help of simulation tools to predict l<strong>and</strong>scape-scale impact of different fire<br />

regimes, <strong>and</strong> to set up <strong>and</strong> to test reliable scenarios in order to limit the impact of fires <strong>and</strong><br />

drought, <strong>and</strong> for helping cork oak conservation. To test the impact of the size of the patches of<br />

cork oak woodl<strong>and</strong>s <strong>and</strong> of different fire regimes on the stability of the oak-shrubl<strong>and</strong> mosaic<br />

we used a simulation approach using data from the field <strong>and</strong> literature to implement <strong>and</strong> to<br />

calibrate the model.<br />

* Corresponding author. Phone: +33 4 42 66 99 24 - fax +33 4 42 66 99 23<br />

Email adress: thomas.curt@cemagref.fr<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Curt & J. Pausas 2010. Are changes in fire regime threatening cork oak-shrubl<strong>and</strong> mosaics<br />

120<br />

2. Methodology<br />

2.1 Study area <strong>and</strong> species<br />

The study area is the Maures massif located in the southeastern part of France (43°3 N, 6.3°E),<br />

which is the largest French area for cork oak (Quercus suber L.). This massif is composed of a<br />

granitic <strong>and</strong> metamorphic basement covered with acidic Cambisols. The climate is typically<br />

Mediterranean <strong>and</strong> subhumid xerothermic. The massif is a hotspot for fires, including zones<br />

with up to five recurrent fires since 1959 (Curt et al., 2009). The Maures massif has features<br />

common to many Mediterranean countries that explain the recurrence of wildfires: the<br />

predominance of human-induced fires owing to population <strong>and</strong> urban growth (Curt <strong>and</strong> Delcros,<br />

2010), the development of large <strong>and</strong> intense summer wildfires during dries <strong>and</strong> windy spells<br />

(Pausas, 2004), enhanced by the abundance of flammable vegetation types (Mouillot et al.,<br />

2003). The cork oak (Quercus suber) populations are protected by the European Union (Habitat<br />

directive 92/43/EEC) due to their high conservation value. They are intermingled with<br />

shrubl<strong>and</strong>s dominated by the resprouter Erica arborea <strong>and</strong> various seeders (Cistus spp.).<br />

2.1 Model <strong>and</strong> simulation plan<br />

In order to simulate the fate of cork oak population, the vegetation dynamics <strong>and</strong> the spread of<br />

fire we used the FATELAND model (Pausas <strong>and</strong> Ramos, 2006), which is a spatially explicit<br />

version of the FATE model (Moore <strong>and</strong> Noble, 1990) implemented under the LASS software<br />

(available at www.uv.es/jgpausas/lass). FATELAND has been used to model the dynamics of<br />

Mediterranean vegetation submitted to different scenarios of disturbance by fires in Spanish<br />

ecosystems (Pausas <strong>and</strong> Lloret, 2007). For all simulations we selected four main species that<br />

correspond to the most representative <strong>and</strong> dominant plant types in the study area, <strong>and</strong> to various<br />

plant functional types: a resprouter tree (cork oak, Quercus suber L.), a resprouter shrub (heath<br />

tree, Erica arborea L.), seeder shrubs (rockroses, Cistus spp.) <strong>and</strong> a perennial grass<br />

(Brachypodium retusum (Pers.) P.Beauv.) that resprouts after fire (Caturla et al., 2000).<br />

We built an artificial l<strong>and</strong>scape of 200 x 200 square cells, each one equivalent to 10 x 10 meters.<br />

The total area is thus 400 ha, i.e. an area sufficient to take into account the spatial interactions<br />

between species, <strong>and</strong> coherent with the maximal distance of seed dispersal. On the basis of our<br />

field survey of vegetation <strong>and</strong> fuels (Curt et al., 2009), we simulated two mosaics of cork oak<br />

woodl<strong>and</strong>s, shrubl<strong>and</strong>s dominated by Erica arborea <strong>and</strong> Cistus species, <strong>and</strong> patches of<br />

Brachypodium grass. The first one corresponded to small patches of cork oak within the<br />

shrubl<strong>and</strong> matrix while the second one corresponded to large patches of cork oak within the<br />

shrubl<strong>and</strong> matrix. Both mosaics had a similar total area for cork oak <strong>and</strong> shrubl<strong>and</strong>s, but they<br />

had different spatial patterning of vegetation.<br />

All simulations had 110 years duration, the last fire being at year 100 to allow vegetation to<br />

recover. In all simulations, the disturbance started at year 10. At the beginning of each<br />

simulation, the initial conditions of vegetation were set equal for each mosaic. During<br />

simulations, each mosaic was submitted to variable fire regimes including different fire<br />

recurrences, fire sizes <strong>and</strong> fire severities:<br />

- the fire recurrence was simulated by setting different fire intervals, i.e. 10, 20, 30, 40<br />

<strong>and</strong> 50 years<br />

- the maximum fire size was set to three levels: small fires (25% of the l<strong>and</strong>scape),<br />

medium fires (50%) <strong>and</strong> large fires (75%)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Curt & J. Pausas 2010. Are changes in fire regime threatening cork oak-shrubl<strong>and</strong> mosaics<br />

121<br />

- the fire severity for cork oak was modeled by two levels of fire response: low (i.e. low<br />

mortality <strong>and</strong> high resprouting rate for cork oak) versus high (high mortality <strong>and</strong> low<br />

resprouting rate for cork oak)<br />

The persistence of all plants was assessed at the end of the simulations using the total<br />

abundance of all cohorts, <strong>and</strong> the abundance of four cohorts representing the life stages (i.e.<br />

seeds, immature individuals, mature trees, <strong>and</strong> mature high trees). In order to analyze the spatial<br />

changes of the mosaics we also compared the spatial patterning of cork oak <strong>and</strong> of the whole<br />

mosaic before <strong>and</strong> after the simulations. For this purpose we assessed the ‘total edge length’<br />

(McGarigal et al., 2002), this statistics being expected to be a good indicator of species’ spatial<br />

interactions (Pausas, 2006). We also tested the spatial autocorrelation for species richness at the<br />

end of simulations by computing the Moran’s I index (Cliff <strong>and</strong> Ord 1981). The comparison of<br />

the effect of different fires regimes on the mosaics was done using multiple analyses of variance<br />

(MANOVA) with the variables describing species abundance, richness <strong>and</strong> patterning as<br />

dependent variables <strong>and</strong> the variables of fire regime as independent variables.<br />

3. Result<br />

The multiple analyses of variance (Table 1) indicated that the overall abundance of cork (i.e. the<br />

number of cells in which cork oak was present whatever the cohort) was quite stable among all<br />

the simulations runs. It varied from 46.5 to 62% of the l<strong>and</strong>scape (mean value 50.7% with a<br />

coefficient of variation of 6.6%) while the area at the beginning of all simulations was 50% of<br />

the l<strong>and</strong>scape. The total abundance of cork oak after simulation did not relate to the type of<br />

mosaic: large <strong>and</strong> small patches of cork oak woodl<strong>and</strong>s resulted in similar cork oak abundance<br />

under a similar fire regime. Likewise, the different cohorts of cork oak (i.e. from seeds to<br />

mature high trees) were not significantly affected by the size of the woodl<strong>and</strong> patches at the<br />

beginning of simulation. Conversely, all the variables of the fire regime clearly impacted the<br />

abundance of cork oak in the l<strong>and</strong>scape (Table 1): its overall abundance decreased with low fire<br />

recurrence <strong>and</strong> small fires. While the presence of cork oak in the l<strong>and</strong>scape was quite stable<br />

among simulations, the different cohorts behaved differently. Seeds were more abundant with a<br />

mean fire interval of 30 to 40 years. Immature individuals were more abundant at a 20-years fire<br />

interval <strong>and</strong> with large fires. Mature cork oak trees tended to increase with longer fire intervals<br />

but not statistically significantly. High mature trees that form the overstory increased at low fire<br />

recurrence, <strong>and</strong> with low severity <strong>and</strong> small fires. In all the simulations, lower disturbance by<br />

fire (i.e. long fire-free intervals, small fires <strong>and</strong> low-severity fires) favored the development of<br />

mature cork oak woodl<strong>and</strong>s dominated by mature trees, with fewer immature individuals.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Curt & J. Pausas 2010. Are changes in fire regime threatening cork oak-shrubl<strong>and</strong> mosaics<br />

122<br />

Table 1. Multiple analysis of variance of cork oak abundance <strong>and</strong> l<strong>and</strong>scape metrics as a<br />

function of the size of the patches of cork oak woodl<strong>and</strong>s <strong>and</strong> the fire regime. The total<br />

abundance of cork oak <strong>and</strong> the abundance of the different cohorts were computed as the mean<br />

value at the end of the simulations (years 100 to 110). NS = non statistically significant (Tukey’s<br />

HSD test 95%)<br />

Size of the Fire Recurrence Fire Size Fire Severity<br />

woodl<strong>and</strong><br />

patches<br />

Cork oak<br />

Total abundance NS 6.59<br />

13.97<br />

NS<br />

P


T. Curt & J. Pausas 2010. Are changes in fire regime threatening cork oak-shrubl<strong>and</strong> mosaics<br />

123<br />

References<br />

Aronson, J., Pereira, J.S., Pausas, J.G., (Eds.), 2009. Cork Oak Woodl<strong>and</strong>s on the Edge: ecology,<br />

Adapative Management, <strong>and</strong> Restoration. Isl<strong>and</strong> Press, Washington, D.C., USA 352 pp.<br />

Baeza, M.J., Raventos, J., Escarre, A., Vallejo, V.R., 2006. Fire risk <strong>and</strong> vegetation structural<br />

dynamics in Mediterranean shrubl<strong>and</strong>. Plant Ecology, 187, 189-201.<br />

Caturla, R.N., Raventós, J., Guàrdia, R., Vallejo, V.R., 2000. Early post-fire regeneration dynamics of<br />

Brachypodium retusum Pers. (Beauv.) in old fields of the Valencia region (eastern Spain). Acta<br />

Oecologica, 21, 1-12.<br />

Curt, T., Adra, W., Borgniet, L., 2009. Fire-driven oak regeneration in French Mediterranean<br />

ecosystems. <strong>Forest</strong> Ecology <strong>and</strong> Management, 258, 2127-2135.<br />

Curt , T., Delcros, P., 2010. Managing road corridors to limit fire hazard. a simulation approach in<br />

southern France. Ecological Engineering, 4, 1-12.<br />

Delitti, W., Ferran, A., Trabaud, L., Vallejo, V.R., 2004. Effects of fire reccurrence in Quercus<br />

coccifera L. shrubl<strong>and</strong>s of the Valencia region (Spain): I. Plant composition <strong>and</strong> productivity.<br />

Plant Ecology, 177, 57-70.<br />

McGarigal, K., Cushman, S.A., Neel, M.C., Ene, E., 2002. FRAGSTATS: spatial pattern analysis<br />

program for categorical maps. University of Massachusetts, Amherst.<br />

Moore, A.D., Noble, I.R., 1990. An Individualistic Model of Vegetation St<strong>and</strong> Dynamics. Journal of<br />

Environmental Management, 31, 61-81.<br />

Moreira, F., Duarte, L., Catry, F., Acacio, V., 2007. Cork extraction as a key factor determining postfire<br />

cork oak survival in a mountain region of southern Portugal. <strong>Forest</strong> Ecology <strong>and</strong><br />

Management, 253, 30-37.<br />

Mouillot, F., Ratte, J.P., Joffre, R., Moreno, J.M., Rambal, S., 2003. Some determinants of the spatiotemporal<br />

fire cycle in a mediterranean l<strong>and</strong>scape (Corsica, France). L<strong>and</strong>scape Ecology, 18,<br />

665-674.<br />

Pausas, J.G., 1997. Resprouting of Quercus suber in NE Spain after fire. Journal of Vegetation<br />

Scienc,e 8, 703-706.<br />

Pausas, J.G., 2004. <strong>Change</strong>s in fire <strong>and</strong> climate in the eastern Iberian Peninsula (Mediterranean basin).<br />

Climatic <strong>Change</strong>, 63, 337-350.<br />

Pausas, J.G., 2006. Simulating Mediterranean l<strong>and</strong>scape pattern <strong>and</strong> vegetation dynamics under<br />

different fire regimes. Plant Ecology ,187, 249-259.<br />

Pausas, J.G., Lloret, F., 2007. Spatial <strong>and</strong> temporal patterns of plant functional types under simulated<br />

fire regimes. International Journal of Wildl<strong>and</strong> Fire, 16, 484-492.<br />

Pausas, J.G., Ramos, J.I., 2006. L<strong>and</strong>scape analysis <strong>and</strong> simulation shell (LASS). Environmental<br />

Modelling & Software, 21, 629-639.<br />

Pons J. & Pausas J.G. 2006. Oak regeneration in heterogeneous l<strong>and</strong>scapes: the case of fragmented<br />

Quercus suber forests in the eastern Iberian Peninsula. <strong>Forest</strong> Ecology & Management 231:<br />

196-204<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A Fleming et al. 2010. <strong>Forest</strong> management <strong>and</strong> climate, through l<strong>and</strong>scape structure, affect the potential for insect outbreak<br />

124<br />

<strong>Forest</strong> management <strong>and</strong> climate, through l<strong>and</strong>scape structure, affect<br />

the potential for insect outbreak<br />

Richard A Fleming 1* , Allan L. Carroll 2 , Jean-Noël C<strong>and</strong>au 1 & Philippe Dreyfus 3<br />

1 Canadian <strong>Forest</strong> Service, Natural Resources Canada, P.O. Box 490, Sault Ste. Marie,<br />

P6A 4C8, Canada<br />

2 Department of <strong>Forest</strong> Sciences, Faculty of <strong>Forest</strong>ry, University of British Columbia,<br />

2424 Main Mall, Vancouver, BC, V6T 1Z4, Canada<br />

3 Institut National de la Recherche Agronomique (INRA), Recherches <strong>Forest</strong>ières<br />

Méditerranéennes (Mediterranean <strong>Forest</strong> Research), Avenue A. Vivaldi, 84000,<br />

Avignon, France<br />

Abstract<br />

The mountain pine beetle (MPB) is endemic to western North American pine forests <strong>and</strong> is<br />

currently causing a very destructive outbreak in western Canada. Historically, expansion further<br />

north <strong>and</strong> east was limited by mountains <strong>and</strong> winter cold, but during recent warm winters the<br />

insect crossed the mountains <strong>and</strong> it now threatens Canada’s eastern forests.<br />

This concern has motivated our study of how l<strong>and</strong>scape structure affects outbreak development<br />

<strong>and</strong> spread. As part of this study, we describe a model of MPB-st<strong>and</strong> dynamics. Based on recent<br />

field work, the model includes a threshold at low density which the local beetle population must<br />

overcome to become eruptive. We show how, from the MPB perspective, l<strong>and</strong>scape structure<br />

responds to climate <strong>and</strong> forest management <strong>and</strong> has strong, indirect affects on the likelihood of<br />

eruption. We discuss the recent under-estimation of MPB spread in Alberta by much more<br />

complex models.<br />

Keywords: mountain pine beetle, st<strong>and</strong> density management diagram, Dendroctonus<br />

ponderosae, forest l<strong>and</strong>scape structure, outbreak spread<br />

1. Introduction<br />

Besides timber, healthy forests provide a variety of non-timber products <strong>and</strong> many ecosystem<br />

services including maintenance of biodiversity, clean water, <strong>and</strong> carbon storage. In Canada’s<br />

forests, change comes primarily in the form of "disturbances". Disturbances occur in many<br />

forms (e.g., storms, fire, insects, disease, <strong>and</strong> logging) <strong>and</strong> over a wide range of scales (Ayres<br />

<strong>and</strong> Lombardero 2000; Dale et al. 2001). Disturbances leave ecological legacies which<br />

determine future species composition, age structure, <strong>and</strong> spatial heterogeneity of the area<br />

(Radeloff et al. 2000) <strong>and</strong> consequently, facilitate or impede the occurrence of future<br />

disturbances (Kulakowski et al. 2003).<br />

Insects are the most diverse class of organisms on earth <strong>and</strong> the major natural cause of depletion<br />

from Canada’s forest productivity. Past outbreaks have engulfed extensive areas (Volney <strong>and</strong><br />

Fleming 2000). Canada’s western forests are now experiencing “the largest insect outbreak in<br />

* Corresponding author.<br />

Email address: rfleming@nrcan.gc.ca<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A Fleming et al. 2010. <strong>Forest</strong> management <strong>and</strong> climate, through l<strong>and</strong>scape structure, affect the potential for insect outbreak<br />

125<br />

Canadian history” (Ono 2004). The mountain pine beetle (Dendroctonus ponderosae) has been<br />

killing mature lodgepole pine over an area extending up to 13 million ha since 1999 (Raffa et al.<br />

2008). The resultant decline in carbon uptake through photosynthesis <strong>and</strong> increase in emissions<br />

from decaying trees has been so great that it has even changed Canada’s western forests from a<br />

small net carbon sink to a large net atmospheric source (Kurz et al. 2008).<br />

Mountain pine beetle (MPB) can successfully attack most western pines, but lodgepole is its<br />

primary host throughout most of its range. Although widespread – occurring from northern<br />

Mexico, through 12 U.S. states <strong>and</strong> 3 Canadian provinces – mountain pine beetle outbreaks in<br />

Canada have been mainly restricted to the southern half of British Columbia <strong>and</strong> the extreme<br />

south-western portion of Alberta. (In 1979, an outbreak occurred in the Cypress Hills at the<br />

southern junction of the Alberta-Saskatchewan border (Ono 2004)). This restriction is not due<br />

to MPB’s host tree. Lodgepole pine extends north into the Yukon <strong>and</strong> Northwest Territories,<br />

<strong>and</strong> east across much of Alberta. Rather, the potential for mountain pine beetle populations to<br />

exp<strong>and</strong> north <strong>and</strong> east has historically been limited by winter cold <strong>and</strong> mountainous terrain.<br />

The substantial shift by mountain pine beetle populations into formerly unsuitable habitats<br />

during the past 30 years as a result of warmer winters has recently enabled the beetle to<br />

overcome the natural barrier of the high mountains (Carroll et al. 2004).<br />

Pre-requisites for a mountain pine beetle outbreak to occur are an abundance of large, mature<br />

pine trees <strong>and</strong> several years of favorable weather. Fire suppression <strong>and</strong> selective harvesting<br />

(for species other than pine) during the latter half of the previous century have more than<br />

tripled the area of mature pine in western Canada. Moreover, mountain pine beetle survival has<br />

increased as a result of warmer winters over much of western Canada, allowing populations to<br />

invade successfully into pine forests in areas that formerly were climatically unsuitable. Thus,<br />

both conditions for an outbreak have coincided, <strong>and</strong> with enough force to produce the largest<br />

MPB outbreak in recorded history. Given the rapid colonization by mountain pine beetles of<br />

areas that formerly were climatically unsuitable in recent decades, continued warming in<br />

western North America associated with climate change will likely allow the beetle to exp<strong>and</strong><br />

its range further northward, eastward <strong>and</strong> toward higher elevations.<br />

In the past, large-scale mountain pine beetle outbreaks collapsed due to localized depletion of<br />

suitable host trees in combination with adverse weather (e.g., an unseasonably cold period or<br />

an extreme winter). The current outbreak may be different. In the absence of unusual weather,<br />

this outbreak may persist by continually moving into new habitats as global warming allows<br />

access to a small, but continual supply of mature pine, thereby maintaining populations at<br />

above-normal levels for some decades into the future (Carroll et al. 2004).<br />

A recent series of benign winters has allowed mountain pine beetle populations to extend their<br />

ranges along the northeastern slopes of the Rockies in Alberta – areas in which the beetle has<br />

not been previously recorded. This has created a very dangerous situation. The MPB is now<br />

approaching the jack pine, Pinus banksiana, forests of northern Alberta <strong>and</strong> Saskatchewan.<br />

There is no known biological barrier to populations of the beetles colonizing jack pine if its<br />

range exp<strong>and</strong>s north to the zone of overlap between jack <strong>and</strong> lodgepole pines. Because jack pine<br />

is susceptible <strong>and</strong> extends through the rest of the boreal all the way to the east coast, the MPB<br />

may be about to gain access to the whole extent of Canada’s boreal forest (Raffa et al. 2008).<br />

Although changes in forest l<strong>and</strong>scape structure caused by selective harvesting <strong>and</strong> fire<br />

suppression are often cited as a contributing factor to the current outbreak, the mechanisms<br />

involved in the interactions between the l<strong>and</strong>scape structure <strong>and</strong> the population dynamics of<br />

mountain pine beetle remain largely unknown. Indeed, the interactions between the spatial<br />

patterns of l<strong>and</strong>scape structures <strong>and</strong> disturbance processes have been a central question in<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A Fleming et al. 2010. <strong>Forest</strong> management <strong>and</strong> climate, through l<strong>and</strong>scape structure, affect the potential for insect outbreak<br />

126<br />

l<strong>and</strong>scape <strong>and</strong> forest ecology. Insect outbreak severity <strong>and</strong> extent have been shown to be<br />

influenced by l<strong>and</strong>scape structure (Coulson et al. 1999; Radeloff et al. 2000; Cairns et al. 2008).<br />

There is a long history of modeling MPB population dynamics (Berryman et al. 1984; Powell et<br />

al. 1996; Logan et al. 1998; Heavilin <strong>and</strong> Powell 2008; Powell <strong>and</strong> Bentz 2009). However, we<br />

still have little idea what triggers a stable, low density, endemic population to erupt to outbreak<br />

levels which can become self-perpetuating through contagious spread between susceptible<br />

st<strong>and</strong>s. Once a population has exceeded the st<strong>and</strong>-level eruptive threshold, its capacity to<br />

contribute to a l<strong>and</strong>scape-scale outbreak will depend on the availability <strong>and</strong> the quality of<br />

susceptible host trees in neighboring st<strong>and</strong>s (Aukema et al. 2006; Safranyik <strong>and</strong> Carroll 2006).<br />

As the resource gets depleted, the l<strong>and</strong>scape structure at a larger scale becomes a critical factor.<br />

2. Methodology<br />

Our over-all objective is to develop a simple, flexible model which can accurately describe the<br />

broad characteristics of the spread of MPB outbreaks in time <strong>and</strong> space over both lodgepole<br />

pine <strong>and</strong> jackpine l<strong>and</strong>scapes. Our goal is to provide a better underst<strong>and</strong>ing of the interactions<br />

between l<strong>and</strong>scape structure <strong>and</strong> the dynamics of mountain beetle populations, particularly<br />

during the eruptive phase. We hope to identify how l<strong>and</strong>scape properties interact with<br />

population dynamics to enable MPB to erupt to outbreak levels <strong>and</strong> the different scales at which<br />

these processes are interacting. The model will also provide a tool to assess the susceptibility of<br />

current <strong>and</strong> future l<strong>and</strong>scapes to mountain pine beetle outbreaks <strong>and</strong> benchmarks for forest<br />

management plans <strong>and</strong> decision support systems.<br />

2.1 Model development<br />

This model can be thought of as comprising 3 interacting components. The site (sub)model<br />

describes the ecological mechanisms that affect the probability that a low density MPB<br />

population can escape from its limiting factors <strong>and</strong> erupt within both a lodgepole pine <strong>and</strong> a<br />

jackpine st<strong>and</strong>. The l<strong>and</strong>scape structure (sub)model describes how pine st<strong>and</strong>s are distributed<br />

over the l<strong>and</strong>scape of concern both in terms of their presence <strong>and</strong> in terms of the ecological<br />

factors which determine the characteristics of their susceptibility to MPB. The dispersal<br />

(sub)model describes how the insects leaving one st<strong>and</strong> are dispersed over the l<strong>and</strong>scape <strong>and</strong><br />

can thus contribute to triggering eruptions in neighboring or more distant st<strong>and</strong>s.<br />

At the forest st<strong>and</strong> level, the dynamics are controlled by the interaction between a host pine<br />

species st<strong>and</strong> model <strong>and</strong> a mountain pine beetle population dynamics model. At the l<strong>and</strong>scape<br />

level, the dynamics of the system emerge from interactions between st<strong>and</strong>s <strong>and</strong> local MPB<br />

populations through insect dispersal. The dynamics of the system also depend on the initial<br />

characteristics of the l<strong>and</strong>scape (i.e. heterogeneity, patch size, connectivity) <strong>and</strong> the evolution of<br />

its structure according to forest management <strong>and</strong> natural disturbance scenarios. Ultimately, the<br />

simulation models will be implemented in CAPSIS, a simulation platform developed at the<br />

Institut National de la Recherche Agronomique (France) to study the dynamics of forest<br />

ecosystems.<br />

3. Results <strong>and</strong> Discussion<br />

This paper focuses on the development of the site (sub)model. Because previous research<br />

(Carroll et al. 2004) showed that the st<strong>and</strong>-level dynamics of MPB populations can be well-<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A Fleming et al. 2010. <strong>Forest</strong> management <strong>and</strong> climate, through l<strong>and</strong>scape structure, affect the potential for insect outbreak<br />

127<br />

described in the context of st<strong>and</strong> density management diagrams (Farnden 1996), it was decided<br />

to use this approach to model the host pine st<strong>and</strong>s.<br />

St<strong>and</strong> density management diagrams (SDMD) are graphical representations of st<strong>and</strong><br />

development that illustrate the interactions between stocking <strong>and</strong> other st<strong>and</strong> parameters such as<br />

mean diameter, top height, <strong>and</strong> volume. In a modeling context, when many st<strong>and</strong>s have to be<br />

“grown” simultaneously, SDMDs present the advantage of eliminating the need to perform<br />

complex mathematical analyses often used in individual tree, distance-dependent growth <strong>and</strong><br />

yield models. This SDMD approach also provides needed flexibility <strong>and</strong> transparency in<br />

preparation for modeling MPB spread into l<strong>and</strong>scapes dominated by a different pine species for<br />

which there has been no prior experience.<br />

The site (sub)model describes the key ecological mechanisms affecting the growth of MPB<br />

populations within even-aged st<strong>and</strong>s of lodgepole pine in the context of SDMDs. This growth is<br />

non-linear: at low densities, tree defences, competitors, <strong>and</strong> lack of suitable habitat severely<br />

restrict MPB populations. But as st<strong>and</strong>s age, the trees become more susceptible to mass attack in<br />

which upwards of a 1000 MPBs will aggregate to overwhelm the defences of an individual tree.<br />

Besides age, other st<strong>and</strong> properties, particularly tree density, also affect a st<strong>and</strong>’s ability to resist<br />

a mass attack. St<strong>and</strong>s are most susceptible beyond 80 years of age, but by at least 160 years the<br />

phloem in their trees has become too thin for a mass-attacking MPB population to successfully<br />

reproduce itself.<br />

In this presentation, we describe the site (sub)model <strong>and</strong> show how it can be used to<br />

infer a variety of different possible patterns of spread for MPB outbreaks. We also show<br />

that the local st<strong>and</strong> dynamics can be crucial determinants of the larger l<strong>and</strong>scape<br />

patterns that emerge when l<strong>and</strong>scape structure <strong>and</strong> dispersal dynamics are also<br />

considered.<br />

References<br />

Aukema, B.H., Carroll, A.L., Zhu, J., Raffa, K.F., Sickley, T.A. <strong>and</strong> Taylor, S.W., 2006.<br />

L<strong>and</strong>scape level analysis of mountain pine beetle in British Columbia, Canada:<br />

spatiotemporal development <strong>and</strong> spatial synchrony within the present outbreak.<br />

Ecography, 29(3): 427-441.<br />

Ayres, M.P. <strong>and</strong> Lombardero, M.J., 2000. Assessing the consequences of global change for<br />

forest disturbance from herbivores <strong>and</strong> pathogens. Science of the Total Environment,<br />

262: 263–286.<br />

Berryman, A.A., Stenseth, N.C. <strong>and</strong> Wollkind, D.J., 1984. Metastability of forest ecosystems<br />

infested by bark beetles. Researches on Population Ecology, 26(1): 13-29.<br />

Cairns, D.M., Lafon, C.W., Waldron, J.D., Tchakerian, M., Coulson, R.N., Klepzig, K.D., Birt,<br />

A.G. <strong>and</strong> Xi, W., 2008. Simulating the reciprocal interaction of forest l<strong>and</strong>scape<br />

structure <strong>and</strong> southern pine beetle herbivory using LANDIS. L<strong>and</strong>scape Ecology, 23(4):<br />

403-415.<br />

Carroll, A. L., Taylor, S.W., Régnière, J. <strong>and</strong> Safranyik, L., 2004. Effects of climate change on<br />

range expansion by the mountain pine beetle in British Columbia. In: T.L. Shore, J.E.<br />

Brooks <strong>and</strong> J.E. Stone (Eds.). Mountain pine beetle symposium: challenges <strong>and</strong><br />

solutions. Natural Resources Canada, Canadian <strong>Forest</strong> Service, Pacific <strong>Forest</strong>ry Centre.<br />

Information Report BC-X-399. Victoria, BC. 298 p.<br />

Coulson, R.N., McFadden, B.A., Pulley, P.E., Lovelady, C.N., Fitzgerald, J.W. <strong>and</strong> Jack, S.B.,<br />

1999. Heterogeneity of forest l<strong>and</strong>scapes <strong>and</strong> the distribution <strong>and</strong> abundance of the<br />

southern pine beetle. <strong>Forest</strong> ecology <strong>and</strong> management, 114(2-3): 471-485.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A Fleming et al. 2010. <strong>Forest</strong> management <strong>and</strong> climate, through l<strong>and</strong>scape structure, affect the potential for insect outbreak<br />

128<br />

Dale, V.H., Joyce, L.A., McNulty, S., Neilson, R.P., Ayres, M.P., Flannigan, M.D., Hanson, P.J.,<br />

Irl<strong>and</strong>, L.C., Lugo, A.E., Peterson, C.J., Simberloff, D., Swanson, F.J., Stocks, B.J. <strong>and</strong><br />

Wotton, B.M., 2001. Climate change <strong>and</strong> forest disturbances. Bioscience, 51(9): 723–<br />

734.<br />

Farnden, C., 1996. St<strong>and</strong> density management diagrams for lodgepole pine, white spruce <strong>and</strong><br />

interior Douglas-fir. Government of Canada, Department of Natural Resources,<br />

Canadian <strong>Forest</strong> Service, Pacific <strong>Forest</strong>ry Centre, Victoria, BC, Inf. Rep. BC-X 360.<br />

Heavilin, J. <strong>and</strong> Powell, J., 2008. A Novel Method of Fitting Spatio-Temporal Models to Data,<br />

with Applications to the Dynamics of Mountain Pine Beetles. Natural Resource<br />

Modeling, 21(4): 489-524.<br />

Kulakowski, D., Veblen, T.T. <strong>and</strong> Bebi, P., 2003. Effects of fire <strong>and</strong> spruce beetle outbreak<br />

legacies on the disturbance regime of a subalpine forest in Colorado. Journal of<br />

Biogeography, 30: 1445-1456.<br />

Kurz, W.A., Dymond, C.C. Stinson, G. Rampley, G.J. Neilson, E.T. Carroll, A.L. Ebata T. <strong>and</strong><br />

Safranyik, L., 2008. Mountain pine beetle <strong>and</strong> forest carbon feedback to climate change.<br />

Nature, 452(7190): 987–990.<br />

Logan, J.A., White, P., et al., 1998. Model analysis of spatial patterns in mountain pine beetle<br />

outbreaks. Theoretical Population Biology, 53(3): 236-255.<br />

Ono, H., 2004. The Mountain Pine Beetle: Scope of the Problem <strong>and</strong> Key Issues in Alberta. In:<br />

T.L. Shore, J.E. Brooks <strong>and</strong> J.E. Stone (Eds.). Mountain pine beetle symposium:<br />

challenges <strong>and</strong> solutions. Natural Resources Canada, Canadian <strong>Forest</strong> Service,<br />

Information Report BC-X-339, Victoria, BC. 298 p.<br />

Powell, J.A. <strong>and</strong> Bentz, B.J., 2009. Connecting phenological predictions with population growth<br />

rates for mountain pine beetle, an outbreak insect. L<strong>and</strong>scape Ecology, 24(5): 657-672.<br />

Powell, J.A., Logan, J.A. <strong>and</strong> Bentz, B.J., 1996. Local projections for a global model of<br />

mountain pine beetle attacks. Journal of Theoretical Biology, 179(3): 243-260.<br />

Radeloff, V.C., Mladenoff, D.J. <strong>and</strong> Boyce, M.S., 2000. Effects of interacting disturbances on<br />

l<strong>and</strong>scape patterns: Budworm defoliation <strong>and</strong> salvage logging. Ecological Applications,<br />

10: 233-247.<br />

Raffa, K.F., Aukema, B.H., Bentz, B.J., Carroll, A.L., Hicke, J.A., Turner, M.G. <strong>and</strong> Romme,<br />

W.H., 2008. Cross-scale drivers of natural disturbances prone to anthropogenic<br />

amplification: the dynamics of bark beetle eruptions. Bioscience, 58(6): 501–517.<br />

Safranyik, L. <strong>and</strong> Carroll, A.L., 2006. The biology <strong>and</strong> epidemiology of the mountain pine<br />

beetle in lodgepole pine forests. In: Les Safranyik <strong>and</strong> Bill Wilson (Eds.). The Mountain<br />

Pine Beetle: A Synthesis of Its Biology, Management <strong>and</strong> Impacts on Lodgepole Pine.<br />

Natural Resources Canada, Canadian <strong>Forest</strong> Service, Pacific <strong>Forest</strong>ry Centre, Victoria,<br />

BC. 3–66.<br />

Volney, W.J.A. <strong>and</strong> Fleming, R.A., 2000. Climate change <strong>and</strong> impacts of boreal forest insects.<br />

Agriculture, Ecosystems <strong>and</strong> Environment, 82(1): 283-294.<br />

Acknowledgements<br />

This work would not be possible without the financial support of Manitoba<br />

Conservation, the Ontario Ministry of Natural Resources, <strong>and</strong> the Saskatchewan<br />

Ministry of Environment through SERG (Spray Efficacy Research Group)-International.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.C. Goebel et al. 2010. How important are riparian forests to aquatic foodwebs in agricultural watersheds<br />

129<br />

How important are riparian forests to aquatic foodwebs in agricultural<br />

watersheds of north-central Ohio, USA<br />

P. Charles Goebel 1* , Charles W. Goss 1 , Virginie Bouchard 2 & Lance R. Williams<br />

3<br />

1<br />

School of Environment <strong>and</strong> Natural Resources, Ohio Agricultural Research <strong>and</strong><br />

Development Center, The Ohio State University, 1680 Madison Ave., Wooster, OH<br />

44691, USA<br />

2 School of Environment <strong>and</strong> Natural Resources, The Ohio State University, 2021<br />

Coffey Rd., Columbus, OH 43210-1085, USA<br />

3<br />

Department of Biology, University of Texas at Tyler, 3900 University Blvd., Tyler,<br />

TX 75799, USA<br />

Abstract<br />

In agricultural l<strong>and</strong>scapes restoring riparian forest patches along streams is a watershed<br />

management priority. There are questions, however, as to the degree to which aquatic food<br />

webs are supported by inputs from small <strong>and</strong> often isolated forested riparian patches in<br />

agricultural l<strong>and</strong>scapes. To examine these contributions we compared the plant communities<br />

<strong>and</strong> stream food webs between forested <strong>and</strong> non-forested riparian patches in an agricultural<br />

l<strong>and</strong>scape <strong>and</strong> used stable isotope analyses to determine whether the primary source of energy<br />

for different levels of aquatic food webs were derived from terrestrial or aquatic sources. We<br />

observed no differences in the δ 13 C signatures of consumers between forested <strong>and</strong> non-forested<br />

riparian areas. Similarly, we also observed few differences in δ 15 N signatures between forested<br />

<strong>and</strong> non-forested sites for different trophic levels. This suggests that there may be other<br />

mechanisms driving the structure of aquatic food webs than basal resources alone.<br />

Keywords: riparian forests, stable isotope analysis, food webs, ecosystem structure<br />

1. Introduction<br />

Riparian forests are dynamic components of the l<strong>and</strong>scape that promote many ecosystem<br />

functions vital to the sustainability <strong>and</strong> productivity of watersheds through their influence on<br />

stream water quality, in-stream habitat, food web structure, <strong>and</strong> ecosystem function (Gregory et<br />

al. 1991; Naiman et al. 1993; Wallace et al. 1997). Although riparian areas are subjected to a<br />

variety of natural disturbances (e.g., flooding, drought, l<strong>and</strong>slides, <strong>and</strong> wildfire) that alter habitat<br />

structure <strong>and</strong> biodiversity (Ilhardt et al. 2000), human disturbances (including agricultural<br />

practices) can alter riparian forests in complex <strong>and</strong> often synergistic ways (Gregory et al. 1991).<br />

In many agricultural l<strong>and</strong>scapes, riparian forests have been removed or greatly modified. As a<br />

result <strong>and</strong> because of the importance of riparian areas to watersheds, riparian corridors are often<br />

protected around streams, <strong>and</strong> they are a major component of most watershed management <strong>and</strong><br />

restoration initiatives.<br />

Under ideal conditions, headwater streams are heterotrophic systems where allochthonous<br />

(carbon) inputs from riparian forest vegetation provides energy <strong>and</strong> nutrients for aquatic food<br />

* Corresponding author. Tel: +1 330-263-2789 – Fax: +1 330-263-3658<br />

Email address: goebel.11@osu.edu<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.C. Goebel et al. 2010. How important are riparian forests to aquatic foodwebs in agricultural watersheds<br />

130<br />

webs, which in turn, provide the assimilative mechanisms for nutrients including nitrogen <strong>and</strong><br />

phosphorus (Vannote et al. 1980; Minshall et al. 1985). In most cases, the processing of<br />

nitrogen, phosphorus, <strong>and</strong> organic matter (carbon) is strongly influenced by channel<br />

morphology (e.g., sinuosity <strong>and</strong> connection to an active floodplain), habitat characteristics (e.g.,<br />

large wood that traps leaves <strong>and</strong> slows downstream movement) <strong>and</strong> diversity <strong>and</strong> structure of<br />

the aquatic community. A diversity of organisms in the aquatic food web sequesters nutrients<br />

<strong>and</strong> slows their downstream transport. In contrast, in managed <strong>and</strong> degraded headwaters, large<br />

proportions of inorganic nutrients enhance the growth of algae, which lowers levels of dissolved<br />

oxygen, consequently limiting the survival of more desirable invertebrate <strong>and</strong> vertebrate<br />

species. The degree to which this assimilative capacity remains intact is particularly critical in<br />

predominantly agricultural watersheds where riparian forests have been removed <strong>and</strong> nutrient<br />

loadings to streams tend to be high. In the absence of organic carbon inputs, as is encountered<br />

when riparian vegetation is removed, the capacity of streams to process <strong>and</strong> retain nutrients can<br />

be lowered significantly (Malanson 1993). In these systems, it remains unclear whether instream<br />

production is an alternative energy source to woody riparian vegetation that structures<br />

aquatic food webs.<br />

In this paper, we examine the relationships between basal resources <strong>and</strong> consumers in small <strong>and</strong><br />

often isolated riparian forest fragments within a larger agricultural l<strong>and</strong>scape. To accomplish<br />

this, we utilize stable isotope analysis to quantify the use of terrestrial <strong>and</strong> aquatic sources of<br />

organic matter by consumers <strong>and</strong> explore how the presence of riparian forest fragments can<br />

affect the source of energy <strong>and</strong> its availability to aquatic consumers in headwater streams. Use<br />

of carbon (δ 13 C) <strong>and</strong> nitrogen (δ 15 N) stable isotopes is a powerful tool to evaluate trophic<br />

relationships in aquatic food webs (Peterson <strong>and</strong> Fry 1987). Because the δ 13 C composition of<br />

animals correspond closely to those of their food sources, it is possible to evaluate the<br />

contribution of various sources of carbon to the energy flux of aquatic systems. Such connection<br />

between food sources <strong>and</strong> the consumers can only be done when sources have distinctive δ 13 C<br />

ratios (Hamilton et al. 1992). It has also been shown that in aquatic systems, allochthonous <strong>and</strong><br />

autochthonous sources of carbon tend to generally have different δ 13 C signatures, providing a<br />

tool to evaluate their incorporation into the aquatic food web. δ 15 N isotope ratios, in contrast,<br />

become more enriched at successive trophic positions.<br />

2. Methodology<br />

2.1 Study area, site selection, <strong>and</strong> field sampling<br />

We conducted our research in the Sugar Creek watershed, located in northeastern Ohio, USA.<br />

The Sugar Creek watershed covers 922 km 2 <strong>and</strong> is dominated by different types of agriculture,<br />

including conventional row-crop production <strong>and</strong> traditional Amish farming practices (Stinner et<br />

al. 1989). The l<strong>and</strong>scape is best described as fragmented, with small isolated riparian forests<br />

<strong>and</strong> woodlots typically comprising less than 10% of the total watershed area. Within this<br />

watershed, we selected nine stream reaches dominated by small forested riparian areas <strong>and</strong> ten<br />

stream reaches dominated by non-forested riparian areas.<br />

At each site, we collected three categories of organic materials: (1) autotrophs from the channel<br />

(i.e., macroalgae, macrophyte, <strong>and</strong> epilithon) <strong>and</strong> the riparian area (i.e., herbaceous vegetation<br />

<strong>and</strong> leaves from trees), (2) detrital material (i.e., coarse particulate organic matter or CPOM,<br />

fine particulate organic matter or FPOM, <strong>and</strong> dissolved organic matter or DOC), <strong>and</strong> (3)<br />

consumers (both macroinvertebrates <strong>and</strong> fishes). Riparian trees were only sampled at forested<br />

sites as they were mostly absent from non-forested sites. All samples were collected in the<br />

spring of 2008 once from each location.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.C. Goebel et al. 2010. How important are riparian forests to aquatic foodwebs in agricultural watersheds<br />

131<br />

For all autotrophs, sampling was conducted by compositing subsamples from different habitats<br />

within the selected reach, resulting in one sample per type of material sampled per site.<br />

Macroalgae <strong>and</strong> macrophytes were r<strong>and</strong>omly sampled through each reach. Macrophytes were<br />

only present at six <strong>and</strong> five of the forested <strong>and</strong> non-forested sites, respectively. Epilithic<br />

microfilm material was collected from cobbles with a brush. In addition to microalgae, these<br />

samples might have contained some detrital material or heterotrophic bacteria. Thus, these<br />

samples are referred to generally as epilithon, whereas the algal component of such sample<br />

would be referred to as epilithic microalgae. In the lab, epilithon sample were filtered onto<br />

precombusted GFF filters <strong>and</strong> invertebrates were removed. Samples of herbaceous species were<br />

collected r<strong>and</strong>omly in the riparian area. At the forested sites, leaves from the dominant three to<br />

four riparian trees were collected. In the laboratory, all material was carefully washed <strong>and</strong> dried.<br />

Three 1-L water samples were collected at the reach <strong>and</strong> brought back to the lab. Coarse<br />

particulate organic matter (> 1mm), fine particulate organic matter (< 1mm, > 0.47μm) <strong>and</strong><br />

dissolved organic matter (


P.C. Goebel et al. 2010. How important are riparian forests to aquatic foodwebs in agricultural watersheds<br />

132<br />

Isotope ratios are expressed as δ 13 C <strong>and</strong> δ 15 N values per mille [‰] relative to the international<br />

reference st<strong>and</strong>ard VPDB using NBS 19, L-SVEC, IAEA-N-1 <strong>and</strong> IAEA-N-2. St<strong>and</strong>ards<br />

deviations of δ 13 C <strong>and</strong> δ 15 N replicate analyses were 0.2 ‰ <strong>and</strong> 0.2 ‰, respectively.<br />

DOC samples were analyzed for δ 13 C using an O.I. Analytical Model 1010 TOC Analyzer (OI<br />

Analytical, College Station, TX) interfaced to a PDZ Europa 20-20 isotope ratio mass<br />

spectrometer (Sercon Ltd., Cheshire, UK). This analysis was conducted at the UC Davis Stable<br />

Isotope Facility, in the Department of Plant Sciences. A 20-mL aliquot of sample was<br />

transferred into a heated digestion vessel <strong>and</strong> reacted sequentially with phosphoric acid then<br />

with sodium persulfate to convert DIC <strong>and</strong> DOC each into a pulse of CO 2 . The two sequential<br />

CO 2 pulses liberated by the chemical treatments are carried in a helium flow to an infra-red gas<br />

analyzer (IRGA), then to the isotope ratio mass spectrometer where the 13 C/ 12 C ratios are<br />

measured <strong>and</strong> compared to ratios of laboratory st<strong>and</strong>ards calibrated against NIST St<strong>and</strong>ard<br />

Reference Materials.<br />

We compared δ 13 C <strong>and</strong> δ 15 N values of dominant basal resources (autotrophs <strong>and</strong> detrital<br />

materials) between forested <strong>and</strong> non-forested streams using a t-test. Likewise, to examine<br />

differences in stable isotope signatures for macroinvertebrate functional feeding groups <strong>and</strong><br />

dominant fish species between forested <strong>and</strong> non-forested riparian areas, we also utilized a t-test.<br />

3. Results<br />

We detected no statistically significant differences in δ 13 C values of dominant autotrophs <strong>and</strong><br />

detrital material between forested <strong>and</strong> non-forested streams (Figure 1A). Algae δ 13 C values were<br />

the most depleted, ranging from -40.5‰ to -22.3‰ while δ 13 C values of the epilithon were the<br />

most enriched (ranging from -23.3‰ to -15.1‰), followed by FPOM δ 13 C values (ranging from<br />

-23.0‰ to -11.3‰). With values ranging collectively from -32.5‰ to -23.6‰, δ 13 C values of<br />

CPOM, riparian vegetation (tree <strong>and</strong> herbaceous), <strong>and</strong> macrophytes were similar (Figure 1A.).<br />

We did detect significant differences in δ 15 N values of basal resources (Figure 1A).<br />

Specifically, we found that macrophytes <strong>and</strong> CPOM were enriched in the non-forested sites<br />

compared to the forested sites (t-test; P = 0.01 <strong>and</strong> 0.01, respectively). No differences in δ 15 N<br />

values of the other autotrophs or detrital material were detected between forests <strong>and</strong> nonforested<br />

riparian areas.<br />

Overall, consumer δ 13 C values across all species are best characterized as enriched, with δ 13 C<br />

values ranging between -28.0‰ <strong>and</strong> -20.2‰ <strong>and</strong> δ 15 N values between 3.2‰ to 8.5‰ for all<br />

macroinvertebrate functional feeding groups (Figure 1). Both filterers <strong>and</strong> grazers had higher<br />

δ 15 N values associated with non-forested riparian areas than forested riparian areas (t-test; P =<br />

0.01 <strong>and</strong> 0.02, respectively). Similarly, δ 13 C values ranged from -22.0‰ to -26.2‰ <strong>and</strong> δ 15 N<br />

values from 10.5‰ to 12.7‰ for selected fish species, with creek chubs having significantly<br />

more enriched δ 13 C values in the non-forested riparian areas (t-test; P = 0.01). These values are<br />

similar to the values for FPOM, CPOM, <strong>and</strong> riparian plants (Figure 1).<br />

4. Discussion<br />

Our results, which were only collected once <strong>and</strong> do not capture seasonal variation in isotopic<br />

signatures, suggest that there are not marked differences in the primary sources of energy<br />

supporting aquatic food webs among forested <strong>and</strong> non-forested headwater streams in this<br />

agricultural watershed. Neither the primary consumers (e.g., grazers) nor higher-level<br />

consumers (e.g., black-nose dace) showed differences in δ 13 C signatures among forested <strong>and</strong><br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.C. Goebel et al. 2010. How important are riparian forests to aquatic foodwebs in agricultural watersheds<br />

133<br />

non-forested reaches. This may be related to the fact that we did not observe clear differences<br />

between δ 13 C signatures for some of the autochthonous resources (epilithon) <strong>and</strong> allochthonous<br />

resources (CPOM <strong>and</strong> riparian plants). It has been observed elsewhere that variation in stream<br />

flows, dissolved CO 2 concentrations, <strong>and</strong> δ 13 C concentrations in dissolved inorganic C can<br />

result in higher than expected algal δ 13 C values that overlap with detrital δ 13 C values which tend<br />

to be more consistent both spatially <strong>and</strong> temporally (Finlay 2004). Consequently, disentangling<br />

energy pathways in this system may require additional study or the use of other stable isotopes<br />

such as hydrogen (δD) (Finlay et al. 2010).<br />

We did observe differences in levels of consumer 15 N enrichment between the forested <strong>and</strong> nonforested<br />

riparian areas. However, we did not observe these differences with the higher-level<br />

consumers as anticipated. We had hypothesized that the increased diversity of basal food web<br />

resources associated with the forested sites would result in δ 15 N signatures of higher-level<br />

consumers. The fact that we did not detect such a relationship suggests a possible linkage with<br />

diet quality <strong>and</strong> stoichiometry associated with the different environments rather than trophic<br />

position. Similar linkages have been suggested by others observing similar patterns in streams<br />

of Arkansas (Dekar et al. 2009). This pattern may also be related to seasonality <strong>and</strong> differences<br />

associated with diet (i.e., more reliance on algae <strong>and</strong> macrophyte sources during the spring as<br />

these basal resources may have higher C:N ratios than detrital sources). However, more<br />

research on these mechanisms is clearly needed to better underst<strong>and</strong> the patterns of 15 N<br />

enrichment in these headwater streams <strong>and</strong> how that may relate to riparian management <strong>and</strong><br />

restoration.<br />

References<br />

Dekar, M.P., Magoulick, D.D., <strong>and</strong> Huxel, G.R. 2009. Shifts in the trophic base of intermittent<br />

stream food webs. Hydrobiologia, 635: 263-277.<br />

Finlay J.C. 2004. Patterns <strong>and</strong> controls of lotic algal stable carbon isotope ratios. Limnology <strong>and</strong><br />

Oceanography, 49: 850-861.<br />

Finlay, J.C., Doucett, R.R., <strong>and</strong> McNeely, C. Tracing energy flow in stream food webs using<br />

stable isotopes of hydrogen. Freshwater Biology. 55: 941-951.<br />

Gregory, S.V., Swanson, F.J., McKee, W.A., <strong>and</strong> Cummins, K.W. 1991. An ecosystem<br />

perspective of riparian zones. Bioscience, 41: 540-551.<br />

Hamilton, S.K., Jr., Lewis, W.M, <strong>and</strong> Sippel, S.J. 1992. Energy sources for aquatic animals in<br />

the Orinoco River floodplain: evidence from stable isotopes. Oecologia, 89: 324-330.<br />

Ilhardt, B.L., Verry, E.S., <strong>and</strong> Palik, Bj.J. 2000. Defining riparian areas. In: E.S. Verry, J.W.<br />

Hornbeck, <strong>and</strong> C.A. Dolloff (Eds.). Riparian Management in <strong>Forest</strong>s of the Continental<br />

Eastern United States. New York, USA: Lewis Publishers: 23-42.<br />

Malanson, G.P. 1993. Riparian l<strong>and</strong>scapes. Cambridge, UK: Cambridge University Press.<br />

Minshall, G.W., K.W. Cummins, T.L. Bott, J.R. Sedell, C.E. Cushing, <strong>and</strong> R.L. Vannote. 1985.<br />

Developments in stream ecosystem dynamics. Ecological Monographs 53:1-25.<br />

Naiman, R.J., Decamps, H., <strong>and</strong> Pollack, M. 1993. The role of riparian corridors in maintaining<br />

regional biodiversity. Ecological Applications, 3: 209-212.<br />

Peterson, B.J., <strong>and</strong> Fry, B. 1987. Stable isotopes in ecosystem studies. Annual Review of<br />

Ecology <strong>and</strong> Systematics, 18: 293–320.<br />

Stinner, D.H., Paoletti, M.G., <strong>and</strong> Stinner, B.R. 1989 Amish agriculture <strong>and</strong> implications for<br />

sustainable agriculture. Agriculture, Ecosystems <strong>and</strong> the Environment, 27: 77-90.<br />

Vannote, R.L., Minshall, G.W., Cummins, K.W., Sedell, J.R., <strong>and</strong> Cushing, C.E. 1980. The<br />

river continuum concept. Canadian Journal of Fisheries <strong>and</strong> Aquatic Sciences, 37: 130-<br />

137.<br />

Wallace, J.B., Eggert, S.L., Meyer, J.L., <strong>and</strong> Webster, J.R. 1997. Multiple trophic levels of a<br />

forest stream linked to terrestrial litter inputs. Science, 77:102-104.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.C. Goebel et al. 2010. How important are riparian forests to aquatic foodwebs in agricultural watersheds<br />

134<br />

A. Basal Resources<br />

18<br />

<strong>Forest</strong>ed Riparian Area<br />

18<br />

Non-<strong>Forest</strong>ed Riparian Area<br />

δ 15 N (‰)<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Algae<br />

Macrophytes<br />

Epilithon<br />

Herbaceous riparian<br />

Woody riparian<br />

CPOM<br />

FPOM<br />

DOC<br />

-40 -35 -30 -25 -20 -15 -10<br />

δ 13 C (‰)<br />

-40 -35 -30 -25 -20 -15 -10<br />

δ 13 C (‰)<br />

B. Consumers<br />

<strong>Forest</strong>ed Riparian Area<br />

18<br />

Non-<strong>Forest</strong>ed Riparian Area<br />

δ 15 N (‰)<br />

15<br />

10<br />

5<br />

0<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Filterers<br />

Gatherers<br />

Grazers<br />

Shredders<br />

Predators<br />

Black-nose dace<br />

Creek chub<br />

White sucker<br />

Central stoneroller<br />

-40 -35 -30 -25 -20 -15 -10<br />

δ 13 C (‰)<br />

-40 -35 -30 -25 -20 -15 -10<br />

δ 13 C (‰)<br />

Figure 1. Stable isotope food web plots for forested <strong>and</strong> non-forested riparian areas, Sugar Creek<br />

watershed, northeastern Ohio for the spring of 2008. Symbols represent the mean δ 13 C <strong>and</strong> δ 15 N<br />

values (± 1 SE) for each basal resource (A) <strong>and</strong> consumer taxonomic grouping (B). Note there were<br />

no woody riparian basal resources sampled for the non-forested riparian areas.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> -New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.R. Iverson et al. 2010. Merger of three modeling approaches to assess potential effects of climate change on trees<br />

135<br />

Merger of three modeling approaches to assess potential effects of climate change<br />

on trees in the eastern United States<br />

Louis R. Iverson 1* , Anantha M. Prasad 1 , Stephen N. Matthews 1,2 & Matthew P. Peters 1<br />

1<br />

Northern Research Station, USDA <strong>Forest</strong> Service, Delaware, OH, USA<br />

2<br />

The Ohio State University, School of Natural Resources, Columbus OH, USA<br />

Abstract<br />

Climate change will likely cause impacts that are species specific <strong>and</strong> significant; modeling is<br />

critical to better underst<strong>and</strong> potential changes in suitable habitat. We use empirical, abundancebased<br />

habitat models utilizing decision tree-based ensemble methods to explore potential<br />

changes of 134 tree species habitats in the eastern United States<br />

(http://www.nrs.fs.fed.us/atlas).To help interpret <strong>and</strong> add value to these outputs, we assigned<br />

<strong>and</strong> calculated Modification Factors for disturbance <strong>and</strong> biological factors that cannot be<br />

specifically assessed with the empirical R<strong>and</strong>om<strong>Forest</strong> approach. We also use a spatially<br />

explicit cellular model, SHIFT, to calculate colonization potentials, based on the abundance of<br />

the species, the distances between occupied <strong>and</strong> unoccupied cells <strong>and</strong> the fragmented nature of<br />

the l<strong>and</strong>scape. By combining results from the three efforts, we are estimating potential impacts<br />

that can be used to aid in management decisions under climate change. These tools are<br />

demonstrated for one species, black oak (Quercus velutina), in northern Wisconsin.<br />

Keywords: Climate <strong>Change</strong>, R<strong>and</strong>om<strong>Forest</strong>, Species Distribution Modeling, Eastern United<br />

States, Trees<br />

1. Introduction<br />

The ranges of tree species in eastern North America have generally shifted northward as the<br />

climate has warmed over the past 14,000+ years since the last ice age (Webb 1992). Evidence is<br />

mounting that tree species, along with many other organisms, are continuing this northward<br />

movement, some at very high rates (Hoegh-Guldberg et al. 2008). There is also increasing<br />

evidence of broad expanses of tree mortality that can be attributed to drier <strong>and</strong> hotter conditions,<br />

often predisposing the forests to insect pest outbreaks (e.g., mountain pine beetle in western<br />

North America) (Allen et al. 2010). Habitats for individual species have, <strong>and</strong> will continue to,<br />

shift independently <strong>and</strong> at different rates, resulting in changing forest community compositions<br />

over time (Webb 1992). Such shifts are likely to occur in the coming decades in the eastern<br />

United States, so that some species will decline in suitable habitat while others will increase to<br />

various degrees. While it is likely that certain habitat will become suitable for some species not<br />

currently present, it is less clear how rapidly – or even whether – those species will migrate into<br />

the region without active human intervention (Higgins <strong>and</strong> Harte 2006). Studies on six eastern<br />

United States species showed that, at the rate of migration typical of the Holocene period (50<br />

km/century in fully forested condition), less than 15% of the newly suitable habitat has even a<br />

remote possibility of being colonized within 100 years (Iverson et al. 2004). The relatively rapid<br />

nature of the projected climate shifts, along with the limits on the rate at which trees can migrate<br />

over a l<strong>and</strong>scape, especially in the current <strong>and</strong> future fragmented state of forests, constrain the<br />

rate of ‘natural’ migration.<br />

* Corresponding author. Tel.:740-368-0097 - Fax:740-368-0152<br />

Email address: liverson@fs.fed.us<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.R. Iverson et al. 2010. Merger of three modeling approaches to assess potential effects of climate change on trees<br />

136<br />

We approach modeling impacts of climate change using a combination of statistical species<br />

distribution models (DISTRIB), literature-based conceptual models (MODFACs), <strong>and</strong> cellbased<br />

spatially explicit models (SHIFT) to quantify colonization probabilities (Fig. 1). The<br />

work presented here is based on modeling the primary individual tree species of the eastern<br />

United States, with results focused on the northern Wisconsin region. We briefly describe the<br />

methods used <strong>and</strong> then present a brief evaluation <strong>and</strong> example of potential changes for one tree<br />

species, black oak (Quercus velutina), along with several other interpretive measures that<br />

complement the models, to assist in further evaluation of vulnerabilities <strong>and</strong> potential<br />

management options.<br />

2. Methodology<br />

2.1 DISTRIB model development<br />

To create the models, current climate variables (1960-1990) are used in statistical model<br />

development, <strong>and</strong> then are swapped with the future climates (~2070-2100) according to several<br />

global circulation models (GCMs) <strong>and</strong> emission scenarios. We used two emission scenarios<br />

developed for the Intergovernmental Panel on Climate <strong>Change</strong> (IPCC): a high level of emissions<br />

that assumes a continued high rate of fossil fuel emissions to 2050 (A1fi), <strong>and</strong> a lower emission<br />

scenario that assumes a rapid conversion to conservation <strong>and</strong> reduced reliance on fossil fuels (B2)<br />

(Nakicenovic <strong>and</strong> al. 2000). The scenarios are also based on the output from two representative<br />

GCMs: the HadleyCM3 model (Pope 2000), <strong>and</strong> the Parallel Climate Model (PCM) (Washington<br />

et al. 2000). We present the HadleyCM3 model projections under the high emissions scenario<br />

(HadHi) as the most extreme warming case in our analysis, <strong>and</strong> the PCM model under the low<br />

emissions scenario (PCMLo) to represent the case with the least warming.<br />

We use the DISTRIB model, a statistical-empirical approach using decision-tree ensembles to<br />

model <strong>and</strong> to predict changes in the distribution of potential habitats for future climates (Fig. 1)<br />

(Prasad et al. 2006), Iverson et al. 2008b). For species information on a total of 134 tree species,<br />

we use the USFS <strong>Forest</strong> Inventory <strong>and</strong> Analysis (FIA) data to build importance value estimates for<br />

each species. We use 38 environmental variables – 7 climate, 9 soil classes, 12 soil characteristics,<br />

5 l<strong>and</strong>scape <strong>and</strong> fragmentation variables, <strong>and</strong> 5 elevation variables – to statistically model current<br />

species abundance with respect to the environment. Because the processes involved are nonlinear<br />

with numerous interactions, we use non-parametric machine-learning approaches using decisiontree<br />

ensembles to predict <strong>and</strong> provide valuable insights into the important predictors influencing<br />

species distributions. Specifically we used a 'tri-model' approach: R<strong>and</strong>om<strong>Forest</strong> for prediction,<br />

bagging trees for assessing the stability among individual decision-trees <strong>and</strong> a single decision tree<br />

to assess the main variables affecting the distribution if the stability among trees proved<br />

satisfactory (Prasad et al. 2006, Iverson et al. 2008). The result is an estimate of the potential<br />

future suitable habitat.<br />

Because models vary in their ability to predict we provide an index of the reliability for each<br />

species. For example, species with highly restricted ranges with low sample size often produce<br />

less satisfactory models as compared to more common species (Schwartz et al. 2006). This<br />

pattern results in quite large differences in the reliability of the predictions among species <strong>and</strong><br />

highlights the need to consider how the model captures the speceis distribution. The 'tri-model'<br />

approach gives us the ability to assess the reliability of the model predictions for each species,<br />

classified as high, medium or low depending on the assessment of the stability of the bagged<br />

trees <strong>and</strong> the R 2 in R<strong>and</strong>om<strong>Forest</strong>. This high rating occurred for 55 of the 134 tree species in our<br />

models. Even if the model reliability was medium or low, R<strong>and</strong>om<strong>Forest</strong> predicts better without<br />

overfitting due to its inherent strengths compared to a single decision-tree (Cutler et al. 2007).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.R. Iverson et al. 2010. Merger of three modeling approaches to assess potential effects of climate change on trees<br />

137<br />

2.2 Modification Factors<br />

The DISTRIB model does not address biological or disturbance factors that may influence a<br />

species’ response to climate change. We conducted a literature assessment of these modifying<br />

factors <strong>and</strong> developed a scoring system to address 9 biological <strong>and</strong> 12 disturbance modification<br />

factors (MODFACs) that influence species distributions. Biological factors we evaluated<br />

include the species’ capacity to regenerate after fire, regenerate vegetatively, establish as<br />

seedlings, disperse, as well as the species’ response to competition for light, elevated CO 2 for<br />

productivity <strong>and</strong> water use efficiency, <strong>and</strong> specificity to specific environmental habitats or<br />

edaphic conditions. For distubance factors, we considered the species’ response to invasive<br />

plants, insects, browse, disease, temperature gradients, fire topkill, wind, ice, pollution, floods,<br />

droughts, <strong>and</strong> harvesting. We also rated their level of uncertainty <strong>and</strong> future relevance under<br />

climate change (e.g., droughts will become more problematic under most future scenarios).<br />

These factors, when considered alongside the speces models, can be used to modify how one<br />

interprets the potential for increasing or decreasing future importance of a species. Each species<br />

is given a set of default scores based on the literature, <strong>and</strong> each factor can be changed by<br />

managers as they consider local conditions. With knowledge of site-specific processes,<br />

managers may be better suited to interpret the models after MODFACs have been considered.<br />

The MODFACs can then be used to modify the interpretation of the potential suitable habitat<br />

models. The goal of this effort is to provide information on the distribution of species under<br />

climate change that accounts for the natural processes that influence the final distribution. In<br />

addition, this approach encourages decision-makers to be actively involved in managing tree<br />

habitats under projected future climatic conditions.<br />

2.3 SHIFT model development<br />

Finally, with SHIFT we are evaluating selected species for their potential migration potentials<br />

from where they exist currently to where, of the new suitable habitat to emerge, they may be<br />

able to colonize over the next 100 years (Fig. 1). SHIFT calculates colonization potentials based<br />

on the abundance of the species, the distances between occupied <strong>and</strong> unoccupied cells, the<br />

quality of the habitat, <strong>and</strong> the fragmented nature of the l<strong>and</strong>scape. The long distance dispersal,<br />

captured by an inverse power function, also depends on stochasticity. By combining output<br />

from DISTRIB <strong>and</strong> SHIFT, we may not only obtain an idea of how the suitable habitat may<br />

move, but also some idea on how far the species may move across the fragmented habitats of<br />

the region. We calibrate movement at the approximate (generous) migration rate of 50 km per<br />

century, according to paleoecological data from the Holocene period (e.g., (Davis 1981). Details<br />

of the method (although under revision now because of newly available computing approaches)<br />

are presented in several publications (Iverson et al. 1999; 2004; Schwartz et al. 2001).<br />

3. Results <strong>and</strong> Discussion<br />

3.1 DISTRIB model<br />

Of 134 species modeled in the eastern United States, we found a total of 73 species of interest<br />

for the region in northern Wisconsin. Using the estimates of potential changes in suitable habitat,<br />

we sorted the species according to their potential to gain, lose, have no change, or enter into the<br />

region from outside. As such, we classify them into 8 vulnerability classes which can be<br />

influenced by climate change classes ranging from most vulnerable to least vulnerable:<br />

Extirpated (Extirp): These species are in northern Wisconsin currently, but all suitable habitat<br />

disappears by 2100. [1 species]; Large Decline (LgDec): Show large declines in suitable habitat ,<br />

especially under the high emissions scenarios [12 species]; Small Decrease (SmDec): Show<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.R. Iverson et al. 2010. Merger of three modeling approaches to assess potential effects of climate change on trees<br />

138<br />

smaller declines, mostly apparent in the high emissions scenarios. [6 species]; No <strong>Change</strong><br />

(NoChg): Show roughly similar suitable habitat now <strong>and</strong> in the future. [6 species]; Small<br />

Increase (SmInc): Have an increased amount of suitable habitat in the future as compared to<br />

current, especially with the higher emissions. [4 species]; Large Increase (LgInc): Have much<br />

higher estimates of suitable habitat in the future as compared to current, especially with the<br />

higher emissions. [17 species]; New Entry Both (NewEntBoth): Have very rarely been currently<br />

detected via FIA sampling in northern Wisconsin, but show potential suitable habitat entering<br />

the region, even under the low emission scenarios. [11 species]; New Entry High (NewEntHi):<br />

Have very rarely been detected via FIA sampling in northern Wisconsin, but show potential<br />

suitable habitat entering the region, especially under higher emissions. [16 species]<br />

We chose to select one example species from class 6, large increaser (LgInc), to illustrate the<br />

process researchers <strong>and</strong> managers may pursue to help chart out a range of mangement choices in<br />

the face of climate change. Our example species, black oak (Quercus velutina), has high model<br />

reliability so we can have higher confidence in the modeled expansion of suitable habitat from<br />

nearly absent to the entire northern Wisconsin range (Fig. 2a), with an increase in suitable<br />

habitat under both low (4-fold increase, Fig. 2b) <strong>and</strong> high (6-fold increase, Fig. 2c) emission<br />

scenarios according to our model.<br />

3.2 Modification Factors<br />

The literature assessment of black oak revealed that it is quite resistant to drought <strong>and</strong> fire<br />

topkill (both projected to increase under climate change), but not so much as compared to many<br />

other oaks. It is moderately affected by diseases (which may also increase) <strong>and</strong> it can succumb<br />

to successive defoliations by gypsy moth. It can regenerate quite well from seed or sprouts with<br />

the likely increased fire under climate change. And it is rather a generalist with respect to<br />

temperatures as well as edaphic <strong>and</strong> environmental habitats, all of which are advantageous for a<br />

species projected to have increased habitat. Overall, both the biological <strong>and</strong> disturbance<br />

modifying factors suggest that the species could do slightly better than modeled from DISTRIB<br />

<strong>and</strong> that it may be well suited for the newly emerging habitats of northern Wisconsin.<br />

3.3 SHIFT model<br />

The preliminary output of 100 repetitive runs of the SHIFT model (1 run = 1% probability of<br />

colonization) for black oak shows an expansion of areas with even a small probability of<br />

colonization of roughly 120-160 km into the new suitable habitat within 100 years (Fig. 2d).<br />

Obviously, variations in colonization probability relate to the current abundance of the species<br />

next to the range boundary (Fig. 2d), <strong>and</strong> the quantity <strong>and</strong> nature of the forest habitat in the<br />

exp<strong>and</strong>ing zone (Fig. 2e). The outputs give us a picture of the relatively small <strong>and</strong> slow<br />

expansion of the species, given the constraints of the current habitat <strong>and</strong> the paleoecologically<br />

derived migration rate of 50 km/century. Of course, should humans intervene to assist in the<br />

migration, the picture could potentially change dramatically.<br />

4. Conclusions<br />

The combination of these three approaches to assessing the likely impacts of climate change<br />

provide a more thorough analysis <strong>and</strong>, we hope, a tool set that managers can begin to use in the<br />

course of their adaptive management decisions in the face of climate change. With DISTRIB,<br />

we provide potential changes in individual species’ suitable habitat under various climate<br />

models <strong>and</strong> scenarios of human responses to this crisis. With the modifying factors, we assess<br />

each species’ capacities <strong>and</strong> vulnerabilities to adapt to various changing conditions <strong>and</strong><br />

disturbances. And with SHIFT, we provide an indication of the rate of ‘natural’ migration<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.R. Iverson et al. 2010. Merger of three modeling approaches to assess potential effects of climate change on trees<br />

139<br />

through the fragmented habitats now existing. We intend to use SHIFT as well to perform<br />

‘experiments’ of l<strong>and</strong>scape manipulation <strong>and</strong> assisted migration to assess these potential<br />

strategies under climate change.<br />

We also show the vast differential in outcomes between a low carbon future (PCMlo) <strong>and</strong> high<br />

carbon future (HADhi) with respect to habitats for one species (it is the case for most species),<br />

<strong>and</strong> thus the critical need for a global effort to reduce carbon emissions.<br />

References<br />

Allen C., Macaladyb A., Chenchounic H., Bachelet D., McDowell N., Vennetier M., Kitzberger<br />

T., Rigling A., Breshear D., Hoggi E., Gonzalezk P., Fensham R., Zhangm Z., Castron<br />

J., Demidavao N., Lim J.-H., Allard G., Running S., Semerci A. <strong>and</strong> Cobb N. 2010. A<br />

global overview of drought <strong>and</strong> heat-induced tree mortality reveals emerging climate<br />

change risks for forests. <strong>Forest</strong> Ecology <strong>and</strong> Management 259: 660-684.<br />

Cutler D.R., Edwards T.C., Beard K.H., Cutler A., Hess K.T., Gibson J. <strong>and</strong> Lawler J.J. 2007.<br />

R<strong>and</strong>om forests for classification in ecology. Ecology 88: 2783-2792.<br />

Davis M.B. 1981. Quaternary history <strong>and</strong> the stability of forest communities. In West D. C. <strong>and</strong><br />

Shugart H. H. (eds.), <strong>Forest</strong> succession: concepts <strong>and</strong> application, pp. 132-153.<br />

Springer-Verlag, New York.<br />

Hoegh-Guldberg O., Hughes L., McIntyre S., Lindenmayer D.B., Parmesan C., Possingham H.P.<br />

<strong>and</strong> Thomas C.D. 2008. Assisted colonization <strong>and</strong> rapid climate change. Science 321:<br />

345 - 346.<br />

Iverson L.R., Prasad A.M., Matthews S.N. <strong>and</strong> Peters M. 2008. Estimating potential habitat for<br />

134 eastern US tree species under six climate scenarios. <strong>Forest</strong> Ecology <strong>and</strong><br />

Management 254: 390-406.<br />

Iverson L.R., Prasad, A.M. <strong>and</strong> Schwartz M.W. 1999. Modeling potential future individual treespecies<br />

distributions in the Eastern United States under a climate change scenario: a<br />

case study with Pinus virginiana. Ecological Modelling 115: 77-93.<br />

Iverson L.R., Schwartz M.W. <strong>and</strong> Prasad A.M. 2004. Potential colonization of new available<br />

tree species habitat under climate change: an analysis for five eastern US species.<br />

L<strong>and</strong>scape Ecology 19: 787-799.<br />

Nakicenovic N. <strong>and</strong> others. 2000. IPCC special report on emissions scenarios. Cambridge<br />

University Press, Cambridge, UK.<br />

Pope V.D. 2000. The impact of new physical parametrizations in the Hadley Centre climate<br />

model -- HadCM3. Climate Dynamics 16, 123-46. 16: 123-146.<br />

Prasad A.M., Iverson L.R. <strong>and</strong> Liaw A. 2006. Newer classification <strong>and</strong> regression tree<br />

techniques: bagging <strong>and</strong> r<strong>and</strong>om forests for ecological prediction. Ecosystems 9: 181-<br />

199.<br />

Schwartz M.W., Iverson L.R. <strong>and</strong> Prasad A.M. 2001. Predicting the potential future distribution<br />

of four tree species in Ohio, USA, using current habitat availability <strong>and</strong> climatic forcing.<br />

Ecosystems 4: 568-581.<br />

Schwartz M.W., Iverson L.R., Prasad A.M., Matthews S.N., O <strong>and</strong> Connor R.J. 2006.<br />

Predicting extinctions as a result of climate change. Ecology 87: 1611-1615.<br />

Washington W.M., Weatherly J.W., Meehl G.A., Semtner Jr. A.J., Bettge, T.W., Craig A.P.,<br />

Str<strong>and</strong> Jr. W.G., Arblaster J.M., Wayl<strong>and</strong> V.B., James R. <strong>and</strong> Zhang Y. 2000. Parallel<br />

climate model (PCM) control <strong>and</strong> transient simulations. Climate Dynamics 16, 755-74.<br />

16: 755-774.<br />

Webb T.I. 1992. Past changes in vegetation <strong>and</strong> climate: lessons for the future. In Peters R. L.<br />

<strong>and</strong> Lovejoy T. E. (eds.), <strong>Global</strong> warming <strong>and</strong> biological diversity, pp. 59-75. Yale<br />

University Press, New Haven, CT.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.R. Iverson et al. 2010. Merger of three modeling approaches to assess potential effects of climate change on trees<br />

140<br />

Fig. 1. Schematic showing approach to modeling potential impacts of climate change on trees <strong>and</strong> birds in<br />

the Eastern United States.<br />

Fig. 2. Potential habitat changes for black oak; a) current distribution; b) future habitat under PCM low<br />

(humans track low carbon emissions); c) future habitat under Had high (humans continue to exp<strong>and</strong><br />

carbon usage); d) SHIFT colonization probability on current distribution; e) forest cover of Wisconsin<br />

region (gray=forest).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.-C. Lin et al. 2010. Immediate effects of typhoon disturbance <strong>and</strong> artificial thinning on understory light environments<br />

141<br />

Immediate effects of typhoon disturbance <strong>and</strong> artificial thinning on<br />

understory light environments in two subtropical forests in Taiwan<br />

Teng-Chiu Lin 1* , Kuo-Chuan Lin 2 , Jeen-Liang Hwong 2 & Hsueh-Fang Wang 3<br />

1<br />

Department of Life Science, National Taiwan Normal University, Taiwan<br />

2<br />

Taiwan <strong>Forest</strong>ry Research Institute, Taiwan<br />

3<br />

Department of Geography, National Taiwan University, Taiwan<br />

Abstract<br />

We compared changes in understory light environments immediately following two typhoons,<br />

<strong>and</strong> artificial thinning (25% <strong>and</strong> 50% of stems removed) in two subtropical forests. Both<br />

typhoon disturbance <strong>and</strong> artificial thinning enhanced understory light availability in Taiwan, but<br />

the enhancement following thinning, 42%, was considerably greater than that following typhoon<br />

disturbance, < 25%. Understory light availability increased 45% <strong>and</strong> 120% after 25% <strong>and</strong> 50%<br />

thinning, respectively. The diffuse nature of canopy disruption following typhoon disturbance<br />

relative to the patchy <strong>and</strong> binary canopy removal associated with artificial thinning was likely<br />

the reason for their very different impacts on understory light environments. It appears that<br />

artificial thinning with more than 25% of stems removed increases understory light availability<br />

to a level that does not naturally occur in low-elevation forests so that forest development<br />

following artificial thinning is likely to be different from that following typhoon disturbance.<br />

Keywords: typhoon disturbance; artificial thinning; understory light environment; Fushan<br />

Experimental <strong>Forest</strong><br />

1. Introduction<br />

The availability of light to understory plants is critical in determining patterns of the understory<br />

plant community (Fahey <strong>and</strong> Puettmann 2007). Because light availability beneath undisturbed<br />

forest canopies is typically low, enhancing understory light availability can increase the growth<br />

<strong>and</strong> survival of both shade-tolerant <strong>and</strong> -intolerant tree seedlings (Chazdon <strong>and</strong> Fetcher 1984;<br />

Oliver <strong>and</strong> Larson 1996). Disturbances that lead to enhancement of understory light availability<br />

create opportunities for seedlings to grow into the mid-canopy <strong>and</strong> overstory <strong>and</strong> therefore play<br />

a key role in determining the structure <strong>and</strong> function of forest ecosystems. Although artificial<br />

thinning may create spatial <strong>and</strong> temporal heterogeneities in the understory light environment<br />

<strong>and</strong> promote the growth of understory plants (Yanai et al. 1998; Wang et al. 2008), the<br />

comparison of its impacts with wind disturbances, to our knowledge, has not been adequately<br />

documented. Such comparisons are critical to evaluate whether artificial thinning results in<br />

patterns <strong>and</strong> processes that are comparable to those following natural disturbances.<br />

Typhoons are the most common natural disturbance in many low-elevation, subtropical<br />

forest ecosystems, such as those in Taiwan with an average of three to four typhoons striking<br />

Taiwan annually (Wu <strong>and</strong> Kuo 1999). Any silvicultural practice that intends to maintain natural<br />

levels of light heterogeneity <strong>and</strong> species diversity in low-elevation forests must consider the role<br />

that typhoon disturbance plays in regulating natural patterns <strong>and</strong> processes of these ecosystems.<br />

* Corresponding author. Tel.: +886-2-7734-6240 - Fax: +886-2-2931-2904<br />

Email address: tclin@ntnu.edu.tw<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.-C. Lin et al. 2010. Immediate effects of typhoon disturbance <strong>and</strong> artificial thinning on understory light environments<br />

142<br />

In 2006, an experimental thinning on a Cryptomeria japonica plantation in central<br />

Taiwan was initiated as an attempt to improve the structural heterogeneity <strong>and</strong> biodiversity of<br />

the forest (Sun 2007). This was the first large <strong>and</strong> comprehensive experimental thinning in<br />

Taiwan, <strong>and</strong> the study included assessment of a wide variety of biotic <strong>and</strong> abiotic consequences<br />

(e.g. vertebrate <strong>and</strong> invertebrate diversity, recruitment of tree species, microclimate,<br />

decomposition, <strong>and</strong> soil respiration) of forest management practices. Although the C. japonica<br />

plantation under investigation is located at a moderate elevation (1500-1700 m), the study<br />

aimed to produce recommendations for isl<strong>and</strong>-wide application (Sun 2007).<br />

The objectives of the current study are to 1) characterize the forest understory light<br />

environment before <strong>and</strong> after the experimental thinning, <strong>and</strong> 2) compare the impacts of artificial<br />

thinning in the C. japonica plantation with those from typhoon disturbances in a mixed<br />

evergreen hardwood forest at Fushan Experimental <strong>Forest</strong> in northeastern Taiwan (Fig. 1). In<br />

both the C. japonica plantation <strong>and</strong> the Fushan Experimental <strong>Forest</strong>, we measured understory<br />

light availability immediately before <strong>and</strong> after artificial thinning <strong>and</strong> typhoon disturbances to<br />

determine whether artificial thinning results in patterns of understory light environments<br />

comparable to those following typhoon disturbances characteristic of many low-elevation<br />

forests in Taiwan.<br />

2. Methodology<br />

2.1 Study site<br />

2.1.1 Cryptomeria japonica plantation: The Zenlun Experimental <strong>Forest</strong><br />

The C. japonica forest plantation is located in central Taiwan between 1500 m <strong>and</strong> 1700<br />

m elevation. The natural vegetation in this area was clear-cut approximately 50 years ago <strong>and</strong><br />

replanted with C. japonica. The mean annual precipitation at the plantation is 3800 mm, <strong>and</strong> the<br />

mean monthly temperature is 17.5 o C (Wang et al. 2007). The mean tree height was<br />

approximately 17 m in 2006 (Chiu 2007 unpublished data).<br />

In 2006, twelve 100 x 100 m plots were established at Zenlun Experimental <strong>Forest</strong><br />

following the design of large forest dynamic plots at the Center for Tropical <strong>Forest</strong> Science of<br />

the Smithsonian Tropical Research Institute (Losos <strong>and</strong> Leigh 2004). The 12 plots were evenly<br />

<strong>and</strong> r<strong>and</strong>omly assigned to three treatments: unthinned, 25% thinning, <strong>and</strong> 50% thinning. Each<br />

plot was divided into a grid of 100 10 x 10 m subplots, which were grouped into 25 20 x 20 m<br />

operating plots. In the 25% thinning, one of the four subplots in each operating plot was<br />

r<strong>and</strong>omly assigned for clear-cutting; two non-adjacent subplots were r<strong>and</strong>omly assigned for<br />

cutting in operating plots assigned a 50% thinning. We established one 100-m transect<br />

perpendicular to the elevational contours at the center of each of the 12 plots. In order to<br />

monitor seasonal variation of understory light environment before thinning, transects at Zenlun<br />

Experimental <strong>Forest</strong> were established one year before thinning. All transects were less than 3 m<br />

from the edges of the thinning subplots; none of our sampling transects ran through the center of<br />

the thinned subplots.<br />

2.1.2 Natural forest: the Fushan Experimental <strong>Forest</strong><br />

The Fushan Experimental <strong>Forest</strong> is a mixed evergreen hardwood forest dominated by<br />

Fagaceae <strong>and</strong> Lauraceae <strong>and</strong> located in northeastern Taiwan at elevations between 500 <strong>and</strong> 1200<br />

m. Typhoons mainly occur between July <strong>and</strong> September, with infrequent typhoons as early as<br />

May <strong>and</strong> as late as December. On average, 1.4 typhoons strike the Fushan Experimental <strong>Forest</strong><br />

each year (Mabry et al. 1998). The mean annual temperature at Fushan is 18.6 o C, <strong>and</strong> the mean<br />

annual relative humidity is 96% (Hsia <strong>and</strong> Hwong 1999). The forest is multistoried with<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.-C. Lin et al. 2010. Immediate effects of typhoon disturbance <strong>and</strong> artificial thinning on understory light environments<br />

143<br />

scattered tree ferns (Alsophila podophylla Hook (Cyatheaceae)) <strong>and</strong> herbaceous cover. We<br />

measured light availability every five meters along a 300-m transect along the eastern ridge of<br />

experimental watershed #1 established in 1994.<br />

2.2 Methods<br />

Understory light environments were investigated using hemispherical photography<br />

(Rich 1990; Lin <strong>and</strong> Chiang 2002). Hemispherical photographs were taken at 1.5 m above the<br />

ground, at 5-m intervals in both forests. Photographs were analyzed to estimate global site<br />

factor (GSF), the proportion of solar radiation reaching a given location, relative to a fully<br />

exposed location with no obstructions (Rich 1990) using Hemiview 2.1 (Delta-T 2000).<br />

Two typhoons, Herb in 1996, <strong>and</strong> Toraji in 2001 were used to evaluate the immediate<br />

effects of typhoon disturbance on understory light environments. According to the Central<br />

Weather Bureau of Taiwan, both typhoon Herb was a strong typhoon (maximum wind velocity<br />

> 51 m.sec-1), <strong>and</strong> typhoon Toraji was a medium-intensity typhoon (maximum wind velocity<br />

33-51 m.sec-1; Table 1).<br />

3. Result<br />

Light indices significantly increased after both typhoon disturbance <strong>and</strong> artificial<br />

thinning (Fig. 1, paired t-test all p-values < 0.001). In the unthinned plots, mean post-thinning<br />

GSF was significantly lower than the mean pre-thinning GSF (paired t-test p < 0.01). The effect<br />

of typhoon disturbance was variable, depending on the strength of the typhoon. The mean<br />

percent changes in understory light indices were significantly greater after typhoon Herb (24%)<br />

than after typhoon Toraji, (7.8%) (Fig. 1, Bonferroni multiple comparisons, both p-values <<br />

0.005).<br />

Artificial thinning caused disproportional increases in understory light indices. The 25%<br />

thinning caused an approximately 42% increase 0.25, while the 50% thinning resulted in an<br />

approximately 120% increase reaching 0.36 (Fig. 2). The increase in understory light indices<br />

was greater after the 25% thinning (approximately 42% enhancement) than after the strong<br />

typhoon Herb (approximately 25% enhancement) (one-way ANOVA, p


T.-C. Lin et al. 2010. Immediate effects of typhoon disturbance <strong>and</strong> artificial thinning on understory light environments<br />

144<br />

4. Discussion<br />

There was a positive relationship between typhoon strength <strong>and</strong> increase of understory<br />

light availabilities following typhoons. Typhoon Herb was a category three typhoon <strong>and</strong> the<br />

strongest to impact Taiwan in the last 3 decades (Longshore 2008). It caused a decrease in the<br />

canopy leaf area index from 2.99 to 2.40 or 20% on the transect (Lin et al. 1999). The 25%<br />

artificial thinning should have also resulted in an approximately 25% removal of canopy leaf<br />

area because all trees located in the thinned subplots were felled. However, the resulting<br />

enhancement of understory light indices (approximately 42%) was significantly greater than that<br />

which followed typhoon Herb (approximately 25%). Note that the magnitude of understory light<br />

enhancement following typhoon Herb was comparable to the magnitude of canopy leaf removal<br />

<strong>and</strong> was likely the result of the spatially distributed effect of typhoon disturbance on canopy<br />

structure. In contrast, thinning removes whole clusters of trees to create gaps of 100m 2 that are<br />

open to the sky <strong>and</strong> resulted in disproportional increases in understory light availabilities which<br />

must be taken into consideration if artificial thinning is used in attempts to enhance understory<br />

light availability to a pre-determined level because the reduction of tree basal area does not<br />

provide a good estimate of increases in canopy transmittance (Hale 2003).<br />

In addition to the disproportional increases in understory light availabilities, the much<br />

greater understory light enhancement following 25% artificial thinning compared to the<br />

strongest typhoon in the last three decades indicates that thinning of 25% or greater is likely to<br />

create understory light environments that do not exist naturally in these forests. The light<br />

availability following artificially thinning, 25%-36% of levels in the open, could lead to a very<br />

different trajectory of forest regeneration.<br />

The very different patterns of post-typhoon GSF for micro-sites that varied in pretyphoon<br />

GSF (Fig. 2) may reflect differences in vulnerability of tree canopies. The increase in<br />

GSF is certainly due to the opening of the forest canopies caused by typhoon disturbance. The<br />

low light micro-sites were likely under taller <strong>and</strong>/or denser tree canopies than the high light<br />

micro-sites. The decrease in GSF after typhoon disturbance in many high-light micro-sites<br />

probably resulted from the larger effect of canopy closure between two repeated measurements<br />

than the effect of typhoon disturbance in these high-light micro-sites. However, typhoons as<br />

intense as Herb in 1996 can completely obliterate this effect of seasonal growth.<br />

After 25% thinning there were still many micro-sites showing decreases in GSF (Fig. 1)<br />

suggesting that the growth potential of micro-sites with high light availability was substantial.<br />

Thus, if such micro-sites are not in or near the thinned subplots, the effect of seasonal growth<br />

cannot be completely offset by thinning by 25% or less.<br />

The very different responses to artificial thinning among micro-sites with similar prethinning<br />

understory light indices (Fig. 2) resulted from their difference in distances relative to<br />

the thinned subplots. Micro-sites near or on the edges of the thinned subplots certainly exhibited<br />

large enhancements of understory light availability. Micro-sites further away were less affected,<br />

<strong>and</strong> the effect of artificial thinning on understory light availability may even be smaller than the<br />

effect of tree growth between the two repeated measurements.<br />

In the natural forest at Fushan Experimental <strong>Forest</strong>, differences in topography <strong>and</strong><br />

species composition are possible causes for the variable impact of typhoon disturbance on<br />

micro-sites of similar pre-typhoon light indices. The disruption to forest canopies <strong>and</strong>, in turn, to<br />

understory light environments were scattered in space resulting in spatially variable impacts on<br />

understory light availability. Thus, the effect of both typhoon disturbance <strong>and</strong> artificial thinning<br />

on understory light environments exhibited large spatial variation.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.-C. Lin et al. 2010. Immediate effects of typhoon disturbance <strong>and</strong> artificial thinning on understory light environments<br />

145<br />

References<br />

Chazdon, R.L. <strong>and</strong> Fetcher, N., 1984. Photographic estimation of photosynthetically active<br />

radiation: evaluation of a computerized technique. Oecologia, 72: 553-564.<br />

Delta-T 2000. HemiView canopy analysis software: user's manual. Version 2.0. Delta-T<br />

Devices Ltd. Cambridge, UK, 79 p.<br />

Fahey, R.T., Puettmann, K.J., 2007. Ground-layer disturbance <strong>and</strong> initial conditions influence<br />

gap partitioning of understorey vegetation. Journal of Ecology, 95: 1098-1109.<br />

Hale, S.E., 2003. The effect of thinning intensity on the below-canopy light environment in a<br />

Sitka spruce plantation. <strong>Forest</strong> Ecology <strong>and</strong> Management, 179: 341-349.<br />

Hsia, Y.J., <strong>and</strong> Hwong, J.L., 1999. Hydrological characteristics of Fu-shan Experimental <strong>Forest</strong>.<br />

Quarterly Journal of Chinese <strong>Forest</strong>ry, 32: 39-51.<br />

Lin, T.C., <strong>and</strong> Chiang, J.M., 2002. Applications of hemispherical photographs in studies of<br />

forest ecology. Taiwan Journal of <strong>Forest</strong> Science, 17: 387-400.<br />

Lin, T.C., Lin, T.T., Chiang, Z.M., Hsia, Y.J., <strong>and</strong> King, H.B., 1999. A study on typhoon<br />

disturbance to the canopy natural hardwood forest in northeastern Taiwan. Quarterly<br />

Journal of Chinese <strong>Forest</strong>ry, 32: 67-78.<br />

Longshore, D., 2008. Encyclopedia of hurricanes, typhoons, <strong>and</strong> cyclones. Checkmark Books,<br />

New York, 468.<br />

Losos, E.C., <strong>and</strong> Leigh Jr., E.G., 2004. Tropical forest diversity <strong>and</strong> dynamism: findings from a<br />

large-scale plot network. The University of Chicago Press, Chicago, 688 p.<br />

Mabry, C.M., Hamburg, S.P. Lin, T.C., Horng, F.W., King, H.B., <strong>and</strong> Hsia, Y.J., 1998.<br />

Typhoon disturbance <strong>and</strong> st<strong>and</strong>-level damage patterns at a subtropical forest in Taiwan.<br />

Biotropica, 30: 238-250.<br />

Oliver, C.D., <strong>and</strong> Larson, B.C., 1996. <strong>Forest</strong> st<strong>and</strong> dynamics. 2nd Edition. John Wiley <strong>and</strong> Sons,<br />

New York, 544 p.<br />

Rich, P.M., 1990. Characterizing plant canopies with hemispherical photographs. Remote<br />

Sensing Review, 5: 13-29.<br />

Sun, I.F., 2007. Managing forest plantation for biodiversity conservation <strong>and</strong> timber production<br />

in Taiwan. In Wang, D.H., <strong>and</strong> Chung, C.H. (Eds) Proceedings of the Symposium on<br />

ecosystem <strong>and</strong> biodiversity conservation on plantation forests in Taiwan <strong>and</strong> France.<br />

Taipei, Taiwan: 1-6.<br />

Wang, D.H., Hwong, J.L., Sun, I.F., Lin, Z.H., Ddeng, S.L., <strong>and</strong> Chung, C.H., 2007. Thinning<br />

effect on st<strong>and</strong> structure of a Sugi Plantation (Cryptomeria Japonica) in Zenlen Area. In<br />

Wang, D.H., <strong>and</strong> Chung, C.H. (Eds) Proceedings of the Symposium on ecosystem <strong>and</strong><br />

biodiversity conservation on plantation forests in Taiwan <strong>and</strong> France. Taipei, Taiwan: 7-<br />

20.<br />

Wang, D.H., Tang, S.C., <strong>and</strong> Liu, C.K., 2008. Four-year monitoring of thinning effects on the<br />

microclimate <strong>and</strong> ground Vegetation in a Taiwania plantation in the Liukuei Experimental<br />

<strong>Forest</strong>, Taiwan. Taiwan Journal of <strong>Forest</strong> Science, 23: 191-198.<br />

Wu, C.C., Kuo, Y.H. 1999. Typhoons affecting Taiwan: current underst<strong>and</strong>ing <strong>and</strong> future<br />

challenges. Bulletin of American Meteorological Society, 80: 67-80.<br />

Yanai, R.D., Twery, M.J., <strong>and</strong> Stout, S.L., 1998. Woody understory response to changes in<br />

overstory density: thinning in Allegheny hardwoods. <strong>Forest</strong> Ecology <strong>and</strong><br />

Management,102: 45-56.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.-C. Lin et al. 2010. Immediate effects of typhoon disturbance <strong>and</strong> artificial thinning on understory light environments<br />

146<br />

Fig. 1. Frequency distribution of light availability before <strong>and</strong> after typhoon disturbance <strong>and</strong><br />

artificial thinning.<br />

Fig. 2. Light availability before <strong>and</strong> after typhoon <strong>and</strong> thinning light availability.<br />

Measurements for each sampling position were sorted based on the rank order of the<br />

beforetyphoon/thinning light measurements.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. L. Martin & P.C. Goebel 2010. Impact of hemlock decline on successional pathways <strong>and</strong> ecosystem function<br />

147<br />

Impact of hemlock decline on successional pathways <strong>and</strong> ecosystem<br />

function at multiple scales in forests of the central Appalachians, USA<br />

Katherine L. Martin * & P. Charles Goebel<br />

School of Environment & Natural Resources, Ohio Agricultural Research &<br />

Development Center, The Ohio State University, 1680 Madison Ave.,<br />

Wooster, OH 44691, USA<br />

Abstract<br />

Hemlock woolly adelgid (HWA) is an invasive, exotic insect causing widespread mortality in<br />

Tsuga canadensis (L.) Carr forests of the eastern United States. T. canadensis is considered a<br />

foundation species that dominates ravine <strong>and</strong> riparian forests across the central <strong>and</strong> southern<br />

Appalachian Mountains. We are working to clarify how the loss of T. canadensis will affect<br />

ecosystem function in forests of the central Appalachians at both local <strong>and</strong> l<strong>and</strong>scape scales.<br />

Using a chronosequence approach, we are examining forests within regions classified as longterm<br />

invaded (> 10 years), recently invaded (5-10 years), <strong>and</strong> intact (not invaded). Initial<br />

analyses indicate hemlock is particularly dominant immediately adjacent to streams, with few<br />

other species in any of the vegetation layers. As this evergreen with poor quality leaf litter<br />

declines, light availability <strong>and</strong> decomposition rates are increasing, changing the successional<br />

pathways of these forests <strong>and</strong> providing resources for additional species including invasive,<br />

non-native plant species.<br />

Keywords: Central Hardwood <strong>Forest</strong>, Ecological disturbance, Hemlock woolly adelgid,<br />

invasive species, Tsuga canadensis<br />

1. Introduction<br />

The forested l<strong>and</strong>scape of eastern North America has been reshaped over the past two centuries<br />

by introduced pests <strong>and</strong> pathogens (Lovett et al. 2006). Tsuga canadensis (L.) Carr has been an<br />

influential canopy tree species throughout these forests from the last major glacial maximum<br />

approximately 10,000 ybp (Allison et al. 1986, Heard <strong>and</strong> Valente 2009). Yet, an invasive<br />

insect may eliminate it from most of its range within the next few decades. Described as a<br />

foundation species (Ellison et al. 2005), the demise of T. canadensis will result in widespread<br />

changes throughout forests of eastern North America. Throughout its range extending from the<br />

southern Appalachian Mountains north to New Engl<strong>and</strong>, the upper Great Lakes <strong>and</strong> eastern<br />

Canada, T. canadensis dominates ecosystem processes. This evergreen conifer contributes a<br />

unique l<strong>and</strong>scape component, adding beta <strong>and</strong> gamma diversity in a matrix of largely deciduous<br />

forests. This may be particularly important in the central <strong>and</strong> southern portions of its range,<br />

where T. canadensis is largely restricted to riparian <strong>and</strong> cove forests along headwater streams.<br />

In the majority of cases, it composes over half of the basal area <strong>and</strong> there is little functional<br />

redundancy with any of the co-occurring deciduous hardwoods.<br />

Introduced in Virginia in the 1950’s, the pest insect hemlock woolly adelgid (Adelges tsugae<br />

Ann<strong>and</strong>; hereafter HWA) has spread from northern Georgia to southern Maine <strong>and</strong> continues to<br />

* Corresponding author. Tel: 01 330 202 3549 - Fax: 01 330 263 3658<br />

Email address: martin.1678@osu.edu<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. L. Martin & P.C. Goebel 2010. Impact of hemlock decline on successional pathways <strong>and</strong> ecosystem function<br />

148<br />

move westward in Pennsylvania (Figure 1). HWA feeds on the parenchyma cells of xylem rays,<br />

which causes the death of branches <strong>and</strong> leads to mortality in T. canadensis (McClure 1991).<br />

There is no evidence that T. canadensis or T. caroliniana Engelm., a species endemic to the<br />

southern Appalachians, have any resistance to HWA. Thus complete mortality occurs in 4-15<br />

years (McClure 1991; Orwig <strong>and</strong> Foster 1998). Evans <strong>and</strong> Gregorie (2007) estimate HWA to<br />

spread at a mean rate of 12 km yr -1 , but both the spread <strong>and</strong> effects are accelerated in southern,<br />

warmer portions of the range (Ford <strong>and</strong> Vose 2007).<br />

On-going T. canadensis mortality is undoubtedly altering forest dynamics including species<br />

composition, diversity, <strong>and</strong> nutrient <strong>and</strong> energy exchanges. T. canadensis forms a foundation<br />

that supports unique species assemblages, including cold-water fishes, birds, <strong>and</strong><br />

macroinvertebrates (Ellison et al. 2005). Across the T. canadensis range, pollen records<br />

following an historic population decline 5,000 ybp provide useful insight into species likely to<br />

respond to hemlock mortality, including Betula, Acer <strong>and</strong> Quercus species (Heard <strong>and</strong> Valente<br />

2009). However, communities that develop following T. canadensis mortality may be regionspecific,<br />

thus creating greater dissimilarity across portions of the l<strong>and</strong>scape once dominated by<br />

hemlock (Ellison et al. 2005; Orwig et al. 2008). Much of the current underst<strong>and</strong>ing of the<br />

impact of T. canadensis mortality is centered in the Northern Hardwoods forest province. Yet, a<br />

large portion of the T. canadensis range lies within the Central Hardwoods forest, a province<br />

extending from West Virginia to Alabama <strong>and</strong> west to Missouri <strong>and</strong> Wisconsin (Fralish 2000).<br />

We are working to clarify the contemporary impact of this introduced disturbance across the<br />

central Appalachians. The objective of our on-going study is to quantify changes in plant<br />

community structure <strong>and</strong> ecosystem function in headwater stream riparian forests of the central<br />

Appalachian Mountains. For the broader study, we are using a chronosequence of time since<br />

HWA invasion to examine the following questions:<br />

1. How will or does T. canadensis mortality impact headwater streams at local scales, moving<br />

from the stream bank upslope<br />

2. How much variation can we expect in these changes both across the central Appalachians<br />

3. How do our results compare to underst<strong>and</strong>ing from the northeastern U.S. <strong>and</strong> southern<br />

Appalachian Mountains<br />

4. What can we predict in terms of forest composition <strong>and</strong> function over time as T. canadensis<br />

mortality progresses <strong>and</strong> forests transition to an alternate ecosystem<br />

Before HWA arrives, we are quantifying the vegetation <strong>and</strong> environmental relationships in T.<br />

canadensis dominated ravines important for natural heritage <strong>and</strong> recreation on the unglaciated<br />

Allegheny Plateau of southeastern Ohio. Therefore, in this paper we focus on our first question:<br />

How will or does T. canadensis mortality impact headwater streams at local scales, moving<br />

from the stream bank upslope According to the projected rate of spread (12 km -1 year, Evans<br />

<strong>and</strong> Gregorie 2007) <strong>and</strong> the current USDA <strong>Forest</strong> Service map of HWA invasion (Figure 1),<br />

Ohio may be invaded within the next few years.<br />

2. Methodology<br />

For the broader study, areas have been identified in the central Appalachian region, including<br />

sites within the Allegheny Mountains of Virginia <strong>and</strong> West Virginia, <strong>and</strong> the Allegheny Plateau<br />

region of eastern Ohio. Using the United States Department of Agriculture (USDA) <strong>Forest</strong><br />

Service HWA distribution map <strong>and</strong> information available from local l<strong>and</strong> managers, we have<br />

identified <strong>and</strong> acquired research permits for study areas in the states of:<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. L. Martin & P.C. Goebel 2010. Impact of hemlock decline on successional pathways <strong>and</strong> ecosystem function<br />

149<br />

1. Virginia (Jefferson <strong>and</strong> George Washington National <strong>Forest</strong>s), where HWA has been present<br />

for 25-30 years <strong>and</strong> hemlocks in invaded st<strong>and</strong>s are likely dead (based on the 4-15 year<br />

mortality timeframe; McClure 1991);<br />

2. West Virginia (New River <strong>and</strong> Gauley River National Recreation Areas, Monongahela<br />

National <strong>Forest</strong>), where HWA has recently invaded (2-5 years) <strong>and</strong> trees are declining; <strong>and</strong>,<br />

3. Ohio (Lake Katherine State Nature Preserve, Sheick Hollow State Nature Preserve, Hocking<br />

State <strong>Forest</strong>) where HWA is not yet present.<br />

In this paper, we focus on data collected in the Unglaciated Allegheny Plateau physiographic<br />

province of southeastern Ohio (Brockman 1998). This province of deep cliffs <strong>and</strong> valleys cut<br />

into the s<strong>and</strong>stone bedrock supports the majority of T. canadensis within Ohio (Black <strong>and</strong> Mack<br />

1976). Three of our steam reaches were located at Lake Katharine, an 817-ha State Nature<br />

Preserve, four were located within the 3,924-ha Hocking State <strong>Forest</strong>, <strong>and</strong> one at Sheick<br />

Hollow, a 61-ha State Nature Preserve. The climate in the area is continental with cold, snowy<br />

winters (mean = 0 o C) <strong>and</strong> warm, humid summers (mean = 21.6 o C) with an average rainfall of<br />

102 cm distributed evenly throughout the year (Kerr 1985; Lemaster <strong>and</strong> Glimore 1989). All<br />

areas were second-growth forest ravines with hemlock approximately 50% or more of the basal<br />

area. Any areas with evidence of recent human disturbance (logging roads, stumps) were<br />

excluded.<br />

We sampled vegetation using a series of five 100-m 2 circular plots (5.64 m radius) centered 10,<br />

30, <strong>and</strong> 50-m from small streams <strong>and</strong> 15 m apart. The species <strong>and</strong> d.b.h. (diameter at breast<br />

height, 1.67 m) of all woody vegetation was recorded. Stems 2.5 - 10 cm d.b.h. were classified<br />

as saplings. All stems > 10 were coded as dominant (full canopy in sun), co-dominant (portions<br />

of canopy in sun), intermediate (top of canopy not overtopped), or sub-canopy (overtopped). For<br />

analyses, the five subplots in each transect were added together for analyses for a 500-m 2 scale.<br />

To underst<strong>and</strong> community characteristics at different slope positions across the ravines, we also<br />

analyzed functional diversity using species guilds developed for the Central Hardwood <strong>Forest</strong> of<br />

the USA by Sutherl<strong>and</strong> et al. (2000). As we are focused on T. canadensis as a foundation<br />

species currently subjected to species-specific mortality, we separated it into its own functional<br />

group for analyses.<br />

Mean species <strong>and</strong> functional richness <strong>and</strong> diversity were compared by transect using one-way<br />

analysis of variance (ANOVA) tests. In cases where means differed, the three transects were<br />

subjected to Tukey’s mean separation test. All analyses were calculated using PASW statistics<br />

18 software (SPSS Inc, Chicago, IL)<br />

3. Results<br />

Our results indicate that in ravines of the Unglaciated Allegheny Plateau of Ohio, T. canadensis<br />

is highly dominant with few other species in the canopy, sub-canopy, or sapling layers (Table<br />

1). Hemlock basal area did not vary by transect, <strong>and</strong> neither did overall basal area. Although<br />

there was not a strong separation (Canopy species richness F = 2.57, P = 0.100, d.f. = 2), Betula<br />

lenta L. <strong>and</strong> Liriodendron tulipifera L., as well as Fagus gr<strong>and</strong>ifolia Ehrh., were more common<br />

adjacent to the stream. Sutherl<strong>and</strong> et al. (2000) categorize B. lenta <strong>and</strong> L. tulipifera as a part of a<br />

long-lived, shade intolerant functional group, indicating they are gap-dependent in hemlock<br />

ravines. On the other h<strong>and</strong>, F. gr<strong>and</strong>ifolia is part of the same extremely shade tolerant, slowgrowing,<br />

<strong>and</strong> long-lived functional group that includes T. canadensis (Sutherl<strong>and</strong> et al. 2000).<br />

Moving from the stream upslope, there was also a slight increase in canopy functional richness<br />

(F= 3.95, P = 0.035, d.f. = 2). This was largely due to increases in Quercus <strong>and</strong> Carya species,<br />

which Sutherl<strong>and</strong> et al. (2000) group into the red oak- hickory (i.e. Quercus rubra L., Q.<br />

coccinea Münchh. <strong>and</strong> Carya spp.) <strong>and</strong> the white oak (Q. alba L., Q. prinus L. ) groups due to<br />

differing germination requirements. Along the 30-m transect, the shade-tolerant Acer-Ulmus<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. L. Martin & P.C. Goebel 2010. Impact of hemlock decline on successional pathways <strong>and</strong> ecosystem function<br />

150<br />

group was also more abundant, largely due A. rubrum L. <strong>and</strong> a few scattered U. rubra Muhl. In<br />

the sub-canopy, a slight increase in species richness (F = 3.87, P = 0.037, d.f. = 2) <strong>and</strong> diversity<br />

(F = 3.10, P = 0.066, d.f.= 2) equated to slightly increased functional richness (F = 3.02, P =<br />

0.070, d.f. = 2). Somewhat mirroring the canopy, B. lenta was more common in sub-canopy at<br />

10 m while A. rubrum <strong>and</strong> Q. prinus were more frequent in the sub-canopy along the 50-m<br />

transect. The sapling layer was the most depauperate with mean species richness below three<br />

species. Saplings were overwhelmingly hemlock <strong>and</strong> were less abundant adjacent to the stream.<br />

Following T. canadensis, A. rubrum <strong>and</strong> F. gr<strong>and</strong>ifolia were the most common species in the<br />

sapling layers, but still at very low numbers (mean below 2 stems per plot).<br />

4. Discussion<br />

On the Unglaciated Allegheny Plateau of southeastern Ohio, T. canadensis dominates steep<br />

ravines formed in s<strong>and</strong>stone bedrock. These ecosystems are characterized by low species <strong>and</strong><br />

functional diversity in the canopy, sub-canopy <strong>and</strong> sapling forest layers. While these forests<br />

remain outside the current HWA invasion, the spread continues <strong>and</strong> will likely reach Ohio<br />

within the next few years. The species that replace T. canadensis will be quite different<br />

functionally. Sutherl<strong>and</strong> et al. (2000) characterize T. canadensis as part of a group including<br />

Acer nigrum Michx. f., A. saccharum Marsh., Carpinus caroliniana Walter, Fagus gr<strong>and</strong>ifolia,<br />

Ostrya virginiana (Mill.) K. Koch , Oxydendrum arboretum (L.) DC, Tilia americana L. <strong>and</strong> T.<br />

heterophylla Vent. These shade tolerant, long-lived species are currently sparse or absent in T.<br />

canadensis ravines. Runkle <strong>and</strong> Whitney (1987) suggest that in this region, poor quality, low<br />

pH soils may be responsible for the lack of these species typically dominant in mesic habitats of<br />

the Central Hardwoods region such as Acer saccharum <strong>and</strong> Tilia spp. This functional group is<br />

also challenged by mortality of F. gr<strong>and</strong>ifolia due to beech bark disease, caused by an<br />

interaction between the pest Cryptococcus fagisuga Ling, <strong>and</strong> the exotic fungi Nectria coccinea<br />

var. faginata Lohman, Watson, <strong>and</strong> Ayers <strong>and</strong> N. galligena Bres. This disease is currently<br />

causing widespread mortality in the northeastern US <strong>and</strong> continues to spread westward.<br />

HWA will shift ravine <strong>and</strong> riparian forests into a different ecosystem state, driven by different<br />

forest functional groups. In Northern Hardwood <strong>Forest</strong>s of the northeastern USA, T. canadensis<br />

is being replaced largely by Betula lenta (Ellison et al. 2005; Orwig et al. 2008). B. lenta is<br />

currently a minor component of T. canadensis ravines in Ohio, however, it is an opportunistic<br />

species well adapted to exploit canopy gaps (Orwig <strong>and</strong> Foster 1998; Catovsky <strong>and</strong> Bazzaz<br />

2002). Ford <strong>and</strong> Vose (2007) also identify Lirodendron tulipifera as a canopy replacement<br />

species in the southern Appalachians. Sutherl<strong>and</strong> et al (2000) group B. lenta, L. tulipifera as<br />

relatively shade-intolerant <strong>and</strong> long-lived. Also included in this group are Juglans nigra <strong>and</strong><br />

Platanus occidentalis which occur infrequently at our plots, but respond well to disturbance.<br />

Our results also indicate that HWA mortality may have differing impacts at local scales. At<br />

upper slope positions, we found a trend of increasing species richness in the sub-canopy <strong>and</strong><br />

canopy, although sapling species richness remained low. This was largely due to increases in<br />

species of Quercus <strong>and</strong> Carya. Species in these groups are responsive to disturbance (Small et<br />

al. 2005) <strong>and</strong> may also exploit gaps created by the death of individual hemlock stems. Thus,<br />

upper slope positions may transition differently than areas immediately adjacent to streams. As<br />

we collect additional data across our mortality chronosequence, these patterns will be refined for<br />

the central Appalachian region.<br />

References<br />

Allison, T.D., Moeller, R.E. <strong>and</strong> Davis, M.B., 1986. Pollen in laminated sediments provides<br />

evidence for a mid-holocene forest pathogen outbreak. Ecology, 67:1101-1105.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. L. Martin & P.C. Goebel 2010. Impact of hemlock decline on successional pathways <strong>and</strong> ecosystem function<br />

151<br />

Black, R.A. <strong>and</strong> Mack, R.N., 1976, Tsuga canadensis in Ohio: Synecological <strong>and</strong><br />

phytogeographic relationships. Plant Ecology, 32: 11-19.<br />

Brockman, C.S., 1998. Physiographic regions of Ohio [map]. Columbus, OH: Division of<br />

Geologic Survey, State of Ohio.<br />

Catovsky, S. <strong>and</strong> Bazzaz. F.A., 2000. The role of resource interactions <strong>and</strong> seedling<br />

regeneration in maintaining a positive feedback in hemlock st<strong>and</strong>s. Journal of Ecology,<br />

88:100-112.<br />

Ellison, A. M., Banks, M. S., Clinton, B. D. , Colburn, E. A., Elliott, K., Ford, C. R. , Foster, D.<br />

R., Kloeppel, B. D. , Knoepp, J. D. , Lovett, G. M., Mohan, J. Orwig, D. A. <strong>and</strong><br />

Rodenhouse, N. L. , 2005. Loss of foundation species: consequences for the structure <strong>and</strong><br />

dynamics of forested ecosystems. Frontiers in Ecology <strong>and</strong> the Environment, 3:479-486.<br />

Evans, A.M. <strong>and</strong> Gregorie, T.G., 2007, A geographically variable model of hemlock woolly<br />

adelgid spread. Biological Invasions, 9:369-382.<br />

Ford, C.R. <strong>and</strong> Vose, J.M., 2007. Tsuga canadensis (L.) Carr. mortality will impact hydrologic<br />

processes in southern Appalachian forest ecosystems. Ecological Applications, 17:1156-<br />

1167.<br />

Fralish, J.S., 2000. The central hardwood forest: Its boundaries <strong>and</strong> physiographic provinces. In:<br />

VanSambeek, J.W., Dawson, J.O., Ponder, F., Loewenstein, E.F., <strong>and</strong> Fralish, J.S. (Eds).<br />

Proceedings of the 13 th Central Hardwood <strong>Forest</strong> Conference. Champaign, IL: USDA<br />

<strong>Forest</strong> Service Northern Research Station: 1-20.<br />

Kerr, J.W., 1985. Soil survey of Jackson County, Ohio. Washington, DC: USDA Soil<br />

Conservation Service. 170 pages.<br />

Lemaster, D.D. <strong>and</strong> Glimore, G.M., 1989. Soil survey of Hocking County, Ohio. Washington,<br />

DC: USDA Soil Conservation Service. 249 pages.<br />

Lovett, G.M., Canham, C.D., Arthur, M.A., Weathers, K.C., <strong>and</strong> Fitzhugh, R.D., 2006. <strong>Forest</strong><br />

ecosystem responses to exotic pests <strong>and</strong> pathogens in eastern North America. BioScience,<br />

56: 395-405.<br />

McClure, M.S., 1991. Density-dependent feedback <strong>and</strong> population cycles in Adelges tsugae<br />

(Homoptera, Adelgidae) on Tsuga canadensis. Environmental Entomology, 20: 258-264.<br />

Orwig, D. A., <strong>and</strong> Foster, D.R., 1998. <strong>Forest</strong> response to the introduced hemlock woolly adelgid<br />

in southern New Engl<strong>and</strong>, USA. Journal of the Torrey Botanical Society, 125 :60–73.<br />

Orwig, D.A., Cobb, R.C. D’Amato, A.W. Kizlinski, M.L., <strong>and</strong> Foster, D.R., 2008. Multi-year<br />

ecosystem response to hemlock woolly adelgid infestation in southern New Engl<strong>and</strong><br />

forests. Canadian Journal of <strong>Forest</strong> Research, 38: 834-843.<br />

Runkle, J.R. <strong>and</strong> Whitney G.G., 1987. Vegetation-site relationships in Lake Katharine State<br />

Nature Preserve Ohio: A northern outlier of the mixed mesophytic forest. Ohio Journal of<br />

Science, 87:36-40.<br />

Small, M.J, Small, C.J. <strong>and</strong> Dreyer, G.D., 2005. <strong>Change</strong>s in a hemlock-dominated forest<br />

following woolly adelgid infestation in southern New Engl<strong>and</strong>. Journal of the Torrey<br />

Botanical Society, 132: 458-470.<br />

Sutherl<strong>and</strong>, E.K., Hale, B.J., <strong>and</strong> Hix, D.M., 2000. Defining species guilds in the Central<br />

Hardwood <strong>Forest</strong>, USA. Plant Ecology, 147: 1-19.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. L. Martin & P.C. Goebel 2010. Impact of hemlock decline on successional pathways <strong>and</strong> ecosystem function<br />

152<br />

Figure 1: USDA <strong>Forest</strong> Service 2009 map of hemlock woolly adelgid spread across the range of T.<br />

canadensis. Study sites have been added with a star, moving west to east: Ohio, West Virginia, <strong>and</strong><br />

Virginia.<br />

Table 1: Comparisons of richness <strong>and</strong> diversity in Tsuga canadensis ravines at three transects centered<br />

upslope at 10, 30, <strong>and</strong> 50 m from headwater streams in Ohio, USA. Comparisons of means were<br />

calculated with one-way analysis of variance. Significant results are indicated: ** α = 0.05, * α = 0.1.<br />

Sub-canopy species richness F = 3.87, P = 0.037; Sub-canopy diversity F = 3.10, P = 0.066; Sub-canopy<br />

functional richness F = 3.02, P = 0.070; Canopy species richness F = 2.57, P = 0.100; Canopy functional<br />

richness F= 3.95, P = 0.035. In all cases d.f.= 2. Means separation as determined by Tukey’s test is<br />

indicated by letters.<br />

10 m 30 m 50 m<br />

Sapling species richness 1.63 ± 0.18 2.5± 0.68 2.38± 0.50<br />

Sapling species diversity 0.20 ± 0.06 0.40 ± 0.19 0.42 ± 0.17<br />

Percent saplings Tsuga canadensis 93.6 ± 2.3 86.6 ± 6.5 77.0 ± 10.4<br />

Sub-canopy species richness** 2.50 ± 0.38 a 3.38 ± 0.38 ab 4.13 ± 0.48 b<br />

Sub-canopy species diversity* 0.48 ± 0.14 a 0.58 ± 0.08 ab 0.93 ± 0.17 b<br />

Sub-canopy functional richness* 2.50 ± 0.48 a 3.25 ± 0.37 ab 3.88 ± 0.13 b<br />

Sub-canopy functional diversity* 0.40 ± 0.14 a 0.45 ± 0.11 ab 0.81 ±0.13 b<br />

Canopy species richness* 4.00 ± 0.27 a 4.63 ± 0.32 ab 5.13 ± 0.44 b<br />

Canopy species diversity 1.22 ± 0.09 1.23 ± 0.10 1.31 ± 0.11<br />

Canopy functional richness** 3.63 ± 0.26 a 4.13 ± 0.23 ab 4.11 ±0.26 b<br />

Canopy functional diversity 1.14 ± 0.09 1.11 ± 0.09 1.11 ± 0.08<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G.M. Pastur et al. 2010. Indirect estimation of l<strong>and</strong>scape uses by Lama guanicoe <strong>and</strong> domestic herbivorous<br />

153<br />

Indirect estimation of l<strong>and</strong>scape uses by Lama guanicoe <strong>and</strong> domestic<br />

herbivorous through the study of diet composition in South Patagonia<br />

Guillermo Martínez Pastur 1* , Rosina Soler Esteban 1 ,<br />

María Vanessa Lencinas 1 & Laura Borrelli 2<br />

1 Centro Austral de Investigaciones Científicas (CADIC CONICET), Argentina<br />

2 Instituto Nacional de Tecnología Agropecuaria (INTA EEA-Bariloche), Argentina<br />

Abstract<br />

Herbivory is central in ecosystem function studies, <strong>and</strong> herbivores influence over regeneration<br />

<strong>and</strong> vegetation composition through browsing. Underst<strong>and</strong>ing diet selection behavior along the<br />

year is important to predict herbivore impacts on regeneration <strong>and</strong> vegetation dynamics under<br />

different scenarios of forest <strong>and</strong> grazing management. The objective was to evaluate diet<br />

composition of herbivorous, <strong>and</strong> correlates it to the use of different vegetation types along the<br />

year. Study was conducted in 100 km² of Tierra del Fuego Isl<strong>and</strong> (Argentina) where 205<br />

floristic surveys were obtained. Feces of native (Lama guanicoe) <strong>and</strong> domestic (Bos taurus <strong>and</strong><br />

Ovis aries) herbivorous were collected. Four laboratory samples per species <strong>and</strong> season were<br />

analyzed, each one composed by five fecal collection. Samples were mixed, dried <strong>and</strong> ground in<br />

a mill (1 mm) for micro-histological analysis. Browsing occurred all around year, where native<br />

<strong>and</strong> domestic herbivorous alternate the uses of different environments. Native species preferred<br />

Nothofagus pumilio forests, except in winter, where domestic species had major predominance.<br />

Beside this, all herbivorous species included grasses in their diet, which were found in open<br />

environments (grassl<strong>and</strong>s <strong>and</strong> peatl<strong>and</strong>s). The knowledge of diet <strong>and</strong> plant distribution at<br />

l<strong>and</strong>scape level allows us to propose management strategies for native <strong>and</strong> domestic<br />

herbivorous.<br />

Keywords: browsing, forest management, fecal micro-histological analysis, Nothofagus,<br />

Argentina.<br />

1. Introduction<br />

Herbivory is central in ecosystem function studies, where large herbivores can produce<br />

important floristic <strong>and</strong> structural changes in forested l<strong>and</strong>scapes, influencing plant productivity<br />

<strong>and</strong> species diversity (Augustine <strong>and</strong> McNaughton 1998; Skarpe <strong>and</strong> Hester 2007). In South<br />

Patagonia, these forested l<strong>and</strong>scapes are mosaics of different site types, where timber-quality<br />

forests rarely constitute large, continuous masses since these are mixed with openl<strong>and</strong>s <strong>and</strong><br />

associated non-timber-quality forest st<strong>and</strong>s (Lencinas et al. 2008), e.g. grassl<strong>and</strong>s, peat-l<strong>and</strong>s,<br />

Nothofagus antarctica forests <strong>and</strong> N. pumilio low quality forests. Beside this, the harvesting of<br />

N. pumilio forests impacts over richness <strong>and</strong> cover of the understory (Martínez Pastur et al.<br />

2002). For this, underst<strong>and</strong>ing diet selection behavior along the seasons is important to predict<br />

herbivore impacts on regeneration <strong>and</strong> vegetation dynamics under different scenarios of forest<br />

<strong>and</strong> grazing management. Lama guanicoe (guanaco) is the only one large native herbivore<br />

(Bonino <strong>and</strong> Fern<strong>and</strong>ez 1994), which compete with domestic herbivores across the l<strong>and</strong>scape<br />

(Bonino <strong>and</strong> Sbriller 1991). The objective was to evaluate diet composition of large herbivorous,<br />

<strong>and</strong> correlate them to the use of different vegetation types along the year.<br />

* Corresponding author. Tel.: +54-2901-422310 int 111- Fax.: +54+2901-430644<br />

Email address: cadicforestal@cadic.gov.ar; cadicforestal@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G.M. Pastur et al. 2010. Indirect estimation of l<strong>and</strong>scape uses by Lama guanicoe <strong>and</strong> domestic herbivorous<br />

154<br />

2. Methodology<br />

The study was conducted in 100 km² of Tierra del Fuego Isl<strong>and</strong> (Argentina) (54°20' SL, 67°52'<br />

WL), where open l<strong>and</strong>s covered 40.0% of the area (grass-l<strong>and</strong>s 36.5% <strong>and</strong> peat-l<strong>and</strong>s 3.5%),<br />

forests covered 59.8% (Nothofagus antarctica forests 19.3% <strong>and</strong> N. pumilio forests 49.5%,<br />

classified as primary unmanaged 26.8%, recently harvested 9.1% <strong>and</strong> old harvesting 13.6%),<br />

<strong>and</strong> lagoons <strong>and</strong> lakes covered 0.2% (Fig. 1). Recently harvested forests (1-5 years) were cutted<br />

using a variable retention method (Martínez Pastur et al. 2009), which include non-harvested<br />

aggregated retention areas (30% of the st<strong>and</strong>) <strong>and</strong> harvested areas with dispersed retention (70%<br />

of the st<strong>and</strong>). Old harvesting (>5 years) was done through selective cuts. Climate is<br />

characterized by short, cool summers <strong>and</strong> long, snowy <strong>and</strong> frozen winters. Mean monthly<br />

temperatures vary from about -7ºC to 14ºC. Absolute temperatures range from -17ºC in July to<br />

22ºC in January. The growing season extends for about 5 months, <strong>and</strong> only 3 months per year<br />

are frost-free. Precipitation is near 400 mm per year, <strong>and</strong> average wind speed is 8 km.h -1 ,<br />

reaching up to 100 km.h -1 during storms (Lencinas et al. 2008). A total of 205 floristic surveys<br />

were obtained recording the cover percentage of vegetation in different site types. Vascular<br />

plants (Dicotyledonae, Monocotyledonae <strong>and</strong> Pteridophytae) were taxonomically classified by<br />

species <strong>and</strong> determined their origin (native or exotic), following Moore (1983) <strong>and</strong> Correa<br />

(1969-1998). Beside this, each species was classified according their relative abundance in<br />

common (more than 0.02% covers in average for the entire census) or rare (less than 0.02%).<br />

Figure 1: Study area, where: yellow = open l<strong>and</strong>s, brown = Nothofagus antarctica forests, green = N.<br />

pumilio forests (dark green = primary unmanaged, green = recently harvested, pale green = old<br />

harvesting), <strong>and</strong> blue = lagoons <strong>and</strong> lakes.<br />

Feces of native (Lama guanicoe) <strong>and</strong> domestic (Bos taurus <strong>and</strong> Ovis aries) herbivorous<br />

were collected along the four seasons (spring, summer, autumn <strong>and</strong> winter). In each sampling,<br />

four areas including different site types were selected, <strong>and</strong> five feces were collected <strong>and</strong> mixed<br />

to complete one pool sample (n = 4 per season per herbivorous species). Pool samples were<br />

oven-dried at 60ºC for 48 h, grounded to


G.M. Pastur et al. 2010. Indirect estimation of l<strong>and</strong>scape uses by Lama guanicoe <strong>and</strong> domestic herbivorous<br />

155<br />

were obtained. Quantification of the species components of the diets was achieved through the<br />

frequencies of each species following Holechek <strong>and</strong> Gross (1982). Plant epidermal fragments<br />

were identified examined using 100x magnification based on their morphological characteristics.<br />

Nothofagus pumilio <strong>and</strong> N. antarctica were identified through the stoma distribution patterns in<br />

the epidermis <strong>and</strong> swellening of cuticle <strong>and</strong> trichomes (Ragonese 1981). Data were analyzed<br />

trough a comparison with a detrended correspondence analysis (DCA) for the environment type<br />

or herbivorous species using plant species values obtained in the different sampling or census<br />

3. Results<br />

Plan diversity (Table 1) greatly changed among the studied vegetation types, <strong>and</strong> large diversity<br />

was shared among the environments (Fig. 2A). Beside this, open environments (mainly<br />

grassl<strong>and</strong>s) presented more exclusive species than the other environments. Harvested forests<br />

increased their diversity with time, including many exotic species (Fig. 2B) compared to the<br />

primary unmanaged forests.<br />

Table 1: Plant richness <strong>and</strong> their codes used for the vegetation type analysis.<br />

Name Code Name Code Name Code<br />

Acaena magellanica ACMA Colobanthus quitensis COQU Pernettya pumila PEPU<br />

Acaena ovalifolia ACOV Coronopus dydimus CODY Phaiophleps biflora PHBI<br />

Acaena pinnatifida ACPI Cotula scariosa COSC Phleum alpinum PHAL<br />

Achillea millefolium ACMI Cystopteris fragilis CYFR Phleum pratense PHPR<br />

Adenocaulon chilense ADCH Dactylis glomerata DAGL Plantago barbata PLBA<br />

Agoseris coronopifolium AGCO Deschampsia antarctica DEAN Poa annua POAN<br />

Agropyron pubiflorum AGPU Deschampsia flexuosa DEFL Poa nemoraris PONE<br />

Agrostis magellanica AGMA Draba magellanica DRMA Poa pratensis POPR<br />

Agrostis perennans AGPE Dysopsis glechomoides DYGL Polygonum aviculare POAV<br />

Agrostis uliginosa AGUL Elymus agropyroides ELAG Pratia longiflora PRLO<br />

Alopecurus magellanicus ALMA Empetrum rubrum EMRU Pratia repens PRRE<br />

Alopecurus pratensis ALPR Epilobium australe EPAU Primula magellanica PRMA<br />

Arenaria serpens ARSE Erigeron myosotis ERMY Ranunculus biternatus RABI<br />

Azorella lycopodioides AZLY Euphrasia antarctica EUAN Ranunculus fuegianum RAFU<br />

Azorella trifurcata AZTR Festuca gracillima FEGR Ranunculus maclovianus RAMA<br />

Berberis buxifolia BEBU Festuca magellanica FEMA Ranunculus uniflorus RAUN<br />

Berberis empetrifolia BEEM Galium antarcticum GAAN Ribes magellanicum RIMA<br />

Blechnum penna-marina BLPE Galium aparine GAAP Rubus geoides RUGE<br />

Bolax gummifera BOGU Galium fuegianum GAFU Rumex acetosella RUAC<br />

Bromus unioloides BRUN Gamochaeta spiciformis GASP Sagina procumbens SAPR<br />

Calamagrostis stricta CAST Gentiana postrata GEPO Schizeilema ranunculus SCRA<br />

Calceolaria biflora CABI Gentianella magellanica GEMG Senecio magellanicus SEMA<br />

Caltha sagitata CASA Geum magellanicum GEMA Senecio vulgaris SEVU<br />

Capsella bursa-pastoris CABU Gunnera magellanica GUMA Stellaria debilis STDE<br />

Cardamine glacialis CAGL Hieracium antarcticum HIAN Stellaria media STME<br />

Carex capitata CACA Holcus lanatus HOLA Taraxacum gillesii TAGI<br />

Carex curta CACU Hordeum comosum HOCO Taraxacum officinale TAOF<br />

Carex decidua CADE Hordeum secalinum HOSE Tetroncium magellanicum TEMA<br />

Carex fuscula CAFU Juncus scheuzerioides JUSC Thlaspi magellanicum THMA<br />

Carex gayana CAGA Leucanthemum vulgare LEVU Trifolium repens TRRE<br />

Carex macloviana CAMA Luzula alopecurus LUAL Triglochin concinna TRCO<br />

Carex magellanica CAMG Lycopodium magellanicum LYMA Triglochin palustris TRPA<br />

Carex sorianoi CASO Nanodea muscosa NAMU Trisetum spicatum TRSP<br />

Carex subantarctica CASU Nothofagus antarctica NOAN Uncinia lechleriana UNLE<br />

Cerastium arvense CEAR Nothofagus pumilio NOPU Veronica serpyllifolia VESE<br />

Cerastium fontanum CEFO Osmorhiza chilensis OSCH Vicia magellanica VIMA<br />

Chiliotrichum diffusum CHDI Osmorhiza depauperata OSDE Viola magellanica VOMA<br />

Cirsium vulgare CIVU Oxalis magellanica OXMA<br />

All the herbivorous mainly grazed in open environments all around the year, <strong>and</strong> were<br />

possible to find the same plant species in their feces analysis (e.g., Carex sp.) in all seasons<br />

(Table 2, Fig. 3). Beside this, in spring, Lama guanicoe consumed also N. pumilio saplings <strong>and</strong><br />

Misodendrum, which were found in forested environments <strong>and</strong> specially in N. pumilio forests,<br />

while domestic herbivorous eaten species of open environments <strong>and</strong> N. antarctica forests too. In<br />

example, N. antarctica saplings, Cotula scariosa or Luzula alopecurus (Fig. 3A).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G.M. Pastur et al. 2010. Indirect estimation of l<strong>and</strong>scape uses by Lama guanicoe <strong>and</strong> domestic herbivorous<br />

156<br />

A<br />

B<br />

Figure 2: DCA for the analysis of plant species distribution: (A) among Nothofagus pumilio forests<br />

(Lenga), N. antarctica forests (Ñire) <strong>and</strong> grassl<strong>and</strong>s <strong>and</strong> peat-l<strong>and</strong>s (Open); (B) among primary<br />

unmanaged N. pumilio forests (PF), recently harvested forests (1-5 years) (RH) <strong>and</strong> old harvesting (>5<br />

years) (OH). Codes for plant species appeared in Table 1.<br />

In spring, while L. guanicoe consumed mainly species of N. pumilio forest, domestic<br />

herbivores incorporated more components of open areas <strong>and</strong> N. antarctica (Fig. 2A). However,<br />

in summer more species were shared among all herbivorous in their diets (Fig. 2B) mainly from<br />

the open environments. In autumn, Lama guanicoe consumed species from N. pumilio forests<br />

<strong>and</strong> open environments, while domestic herbivorous eaten species of open environments <strong>and</strong> N.<br />

antarctica forests (Fig. 3C).<br />

A<br />

B<br />

C<br />

D<br />

Figure 3: DCA for the analysis of diet feces composition of guanacos (G), cows (C) <strong>and</strong> sheeps (S)<br />

according the seasonal sampling (A = spring, B = summer, C = autumn, D = winter). Codes appeared in<br />

Table 2.<br />

Finally, during winter (Fig. 3D), domestic herbivorous consumed plants from N.<br />

pumilio forest <strong>and</strong> open environment species (e.g., N. pumilio saplings, Plantago barbata <strong>and</strong><br />

Trisetum spicatum), while L. guanicoe preferred species from open environments (e.g., Senecio<br />

magellanicus or N. antarctica saplings). Thus, native <strong>and</strong> domestic herbivorous alternated the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G.M. Pastur et al. 2010. Indirect estimation of l<strong>and</strong>scape uses by Lama guanicoe <strong>and</strong> domestic herbivorous<br />

157<br />

uses of the different environments around the year, where an incompatibility of seasonal uses<br />

was observed.<br />

Table 2: Annual percentage of plant genus or species registered in diets, <strong>and</strong> their codes used for the<br />

analysis.<br />

Species Code G C S<br />

Acaena AC 1.6% 1.2% 2.9%<br />

Achillea millefolium ACMI 0.0% 0.1% 1.2%<br />

Agrostis AG 5.7% 8.7% 10.0%<br />

Alopecurus magellanicus ALMA 3.1% 2.9% 3.7%<br />

Arjona AR 0.1% 0.0% 0.2%<br />

Berberis BE 3.1% 0.5% 1.1%<br />

Blechnum penna-marina BLPE 0.9% 1.6% 1.5%<br />

Bromus unioloides BRUN 0.1% 0.0% 0.4%<br />

Calceolaria biflora CABI 0.1% 0.0% 0.0%<br />

Capsella bursa-pastoris CABU 0.0% 0.0% 0.0%<br />

Carex CA 16.8% 29.3% 20.5%<br />

Cerastium CE 2.3% 4.2% 3.3%<br />

Cotula scariosa COSC 0.7% 0.5% 2.3%<br />

Deschampsia DE 9.0% 8.1% 7.4%<br />

Empetrum rubrum EMRU 2.8% 1.7% 1.3%<br />

Epilobium australe EPAU 0.3% 0.0% 0.5%<br />

Erigeron myosotis ERMY 0.2% 0.0% 0.1%<br />

Erodium cicutarium ERCI 0.1% 0.0% 0.2%<br />

Festuca magellanica FEMA 3.6% 9.6% 7.2%<br />

Galium GA 1.2% 0.4% 0.5%<br />

Geum magellanicum GEMA 2.9% 0.0% 0.0%<br />

Gunnera magellanica GUMA 0.7% 0.3% 0.5%<br />

Hordeum HO 0.5% 0.0% 0.0%<br />

Juncus JU 2.6% 2.0% 4.4%<br />

Luzula alopecurus LUAL 0.2% 2.1% 1.2%<br />

Misodendron MI 10.7% 3.6% 5.7%<br />

Nanodea muscosa NAMU 0.1% 0.2% 0.1%<br />

Nothofagus antarctica NOAN 2.7% 1.6% 3.8%<br />

Nothofagus pumilio NOPU 19.1% 10.1% 7.3%<br />

Osmorhiza OS 0.3% 0.0% 0.6%<br />

Pernettya PE 0.6% 0.5% 1.8%<br />

Plantago PL 2.1% 1.4% 0.5%<br />

Ranunculus RA 0.3% 0.0% 0.1%<br />

Rubus geoides RUGE 1.3% 0.0% 1.1%<br />

Rumex RU 0.1% 0.0% 0.1%<br />

Senecio allophyllus SEAL 0.6% 0.1% 0.0%<br />

Senecio magellanicus SEMA 1.1% 0.3% 2.4%<br />

Trifolium repens TRRE 0.0% 0.0% 0.4%<br />

Trisetum spicatum TRSP 0.4% 0.9% 1.1%<br />

Uncinia lechleriana UNLE 0.0% 0.2% 0.2%<br />

Veronica VE 0.0% 0.0% 0.1%<br />

Vicia VI 0.2% 0.0% 0.5%<br />

Sphagnum SP 1.9% 7.8% 3.6%<br />

4. Discussion<br />

Lama guanicoe is a generalist species, which can adapt to a wide spectrum of environments<br />

(Raedeke 1980; Puig et al. 1997), from closed forests where mainly consumed leaves <strong>and</strong><br />

sprouts of Nothofagus saplings (Martínez Pastur et al. 1999; Pulido et al. 2000) to sub-alpine<br />

grassl<strong>and</strong>s (Rebertus et al. 1997), where grasses <strong>and</strong> sedges are the main food. Domestic<br />

herbivorous mostly include grasses in their diet, but also can consumed saplings in adverse<br />

conditions, mainly in winter when forest are preferred among other site types because provides<br />

them with protection. Lama guanicoe avoid presence of domestic herbivorous, using<br />

environments free of them, independently of the season. This native species is a natural<br />

component of these native forests, but the impact in the unmanaged <strong>and</strong> harvested st<strong>and</strong>s can<br />

increase according to the livestock density in the area. The knowledge of diet <strong>and</strong> plant<br />

distribution at l<strong>and</strong>scape level allows us to propose management strategies for native <strong>and</strong><br />

domestic herbivorous, minimizing the impact to the forestry activities.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G.M. Pastur et al. 2010. Indirect estimation of l<strong>and</strong>scape uses by Lama guanicoe <strong>and</strong> domestic herbivorous<br />

158<br />

References<br />

Augustine, D.J. <strong>and</strong> McNaughton, S.J., 1998. Ungulate effects on functional species<br />

composition of plant communities: herbivore selectivity <strong>and</strong> plant tolerance. J. Wildlife<br />

Manage., 62: 1165-1183.<br />

Bonino, N. <strong>and</strong> Pelliza Sbriller, A., 1991. Comparación de las dietas del guanaco, ovino y<br />

bovino en Tierra del Fuego, Argentina. Turrialba 41(8): 452-457.<br />

Bonino, N. <strong>and</strong> Fernández, E., 1994. Distribución general y abundancia relativa de guanacos<br />

(Lama guanicoe) en diferentes ambientes de Tierra del Fuego, Argentina. Ecología<br />

Austral, 4(2): 79-85.<br />

Correa, M.N., 1969-1998. Flora Patagónica. Colección Científica INTA Tomo 8. Parts II, III,<br />

IVb, V, VI y VII. Buenos Aires.<br />

Holechek, J.L. <strong>and</strong> Gross, B.D., 1982. Evaluation of different calculation procedures for<br />

microhistological analysis. J. Range Manage., 36: 721-723.<br />

Latour, M. <strong>and</strong> Pelliza Sbriller, A., 1981. Clave para la determinación de la dieta de herbívoros<br />

en el noroeste de la Patagonia. Rev. Investigación Agrícola (INTA), 16: 109-157.<br />

Lencinas, M.V., Martínez Pastur, G., Rivero, P. <strong>and</strong> Busso, C., 2008. Conservation value of<br />

timber quality vs. associated non-timber quality st<strong>and</strong>s for understory diversity in<br />

Nothofagus forests. Biodiv. Conserv., 17: 2579-2597.<br />

Martínez Pastur, G., Peri, P.L., Fernández, C., Staffieri, G. <strong>and</strong> Rodriguez, D., 1999. Desarrollo<br />

de la regeneración a lo largo del ciclo del manejo forestal de un bosque de Nothofagus<br />

pumilio: 2. Incidencia del ramoneo de Lama guanicoe. Bosque, 20(2): 47-53.<br />

Martínez Pastur, G., Peri, P.L., Fernández, M.C., Staffieri, G. <strong>and</strong> Lencinas, M.V., 2002.<br />

<strong>Change</strong>s in understory species diversity during the Nothofagus pumilio forest<br />

management cycle. J. For. Res., 7: 165-174.<br />

Martínez Pastur, G., Lencinas. M.V., Cellini, J.M., Peri, P.L. <strong>and</strong> Soler Esteban, R., 2009.<br />

Timber management with variable retention in Nothofagus pumilio forests of southern<br />

Patagonia. For. Ecol. Manage., 258: 436-443.<br />

Moore, D., 1983. Flora of Tierra del Fuego. Anthony Nelson, Missouri Botanical Garden, UK.<br />

395 pp.<br />

Puig, S., Videla, F. <strong>and</strong> Cona, M.I., 1997. Diet <strong>and</strong> abundance of the guanaco (Lama guanicoe<br />

Müller 1776) in four habitats of northern Patagonia, Argentina. J. Arid Environ., 36: 343-<br />

357.<br />

Pulido, F., Díaz, B. <strong>and</strong> Martínez Pastur, G., 2000. Incidencia del ramoneo del guanaco (Lama<br />

guanicoe) sobre la regeneración de lenga (Nothofagus pumilio) en bosques de Tierra del<br />

Fuego, Argentina. Investigación Agraria: Sistemas y Recursos <strong>Forest</strong>ales, 9(2): 381-394.<br />

Raedake, K., 1980. Food habitats of the guanaco (Lama guanicoe) of Tierra del Fuego, Chile.<br />

Turrialba, 30: 177-181.<br />

Ragonese, A.M., 1981. Anatomía foliar de especies sudamericanas de Nothofagus Bl.<br />

(Fagaceae). Darwiniana, 23(2-4): 587-603.<br />

Rebertus, A., Kitzberger, T., Veblen, T. <strong>and</strong> Roovers, L., 1997. Blowdown history <strong>and</strong><br />

l<strong>and</strong>scape patterns in the Andes of Tierra del Fuego, Argentina. Ecology, 78(3): 678-692.<br />

Skarpe, C. <strong>and</strong> Hester, A.J., 2007. The impacts of browsing <strong>and</strong> grazing on plant population<br />

dynamics. In: Gordon, I. <strong>and</strong> H. Prins (Eds.). Ecology of grazing <strong>and</strong> browsing of<br />

mammalian herbivores. Springer Verlag. pp. 217-262.<br />

Sepúlveda, L., Pelliza, A. <strong>and</strong> Manacorda, M., 2004. Importancia de los tejidos no epidérmicos<br />

en el microanálisis de dieta. Ecología Austral, 14: 31-38.<br />

Sparks, D., <strong>and</strong> Malechek, J.C., 1968. Estimating percentage dry weigth in diets using a<br />

microscopic technique. J. Range Manage., 21: 264-265.<br />

Williams, O., 1969. An improved technique for identification of plants fragments in herbivore<br />

feces. J. Range Manage., 22(2): 51-52.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.W Saragih et al. 2010. Effects of endozoochorous seed dispersal on the soil seed bank <strong>and</strong> vegetation in the woodl<strong>and</strong> area<br />

159<br />

Effects of endozoochorous seed dispersal on the soil seed bank <strong>and</strong><br />

vegetation in the woodl<strong>and</strong> area<br />

Evi Warintan Saragih 1 , Jan Bokdam 2 & Wim Braakhekke 2<br />

1 Saragih,. Animal Husb<strong>and</strong>ry Faculty. Papua University. Manokwari, Indonesia<br />

2 Bokkdam, Jan <strong>and</strong> 2 Braakhekke, Wim. Department of Nature Conservation <strong>and</strong> Plant<br />

Ecology Wageningen University, The Netherl<strong>and</strong>s<br />

Abstract<br />

In the framework of nature conservation <strong>and</strong> restoration, some experts found the potential of<br />

endozoochorous seed dispersal in a semi-natural l<strong>and</strong>scape. Predicting seed input by large<br />

herbivores through germination test of seeds in dung is needed to figure out the role of large<br />

herbivores. The contribution of large herbivores to ecological restoration sites can be<br />

determined by gathering quantitative information of the dung seed content <strong>and</strong> compared it with<br />

soil seed bank content <strong>and</strong> aboveground vegetation in grazed <strong>and</strong> ungrazed woodl<strong>and</strong> area. The<br />

seed density <strong>and</strong> species richness of two areas as well as dung content were evaluated by<br />

germination test in the green house to find out effect of grazing regime. The result show that<br />

cattle grazing had a positive effect on distribution of vegetation from lawn area to woodl<strong>and</strong><br />

area. Species richness in graze area is higher than ungrazed one. Grazing reduced the cover of<br />

grazing sensitive <strong>and</strong> transport exclusive species to grazed woodl<strong>and</strong> area. Grazing affected the<br />

number of seeds in the soil seed bank sample in the woodl<strong>and</strong> area by creating gaps, stimulating<br />

losses by germination. In conclusions, seed density <strong>and</strong> species richness in the soil seed bank<br />

are less directly affected by large herbivore but direct effect on above ground vegetation.<br />

Keywords: endozoochorous, seed dispersal, herbivores, vegetation<br />

1. Introduction<br />

Seed dispersal can be defined as the movement of seeds from one place to another by agents<br />

including wind, water <strong>and</strong> animals. Seed dispersal is a bottleneck for vegetation development<br />

<strong>and</strong> restoration of isolated (semi) natural relict of nature conservation areas, particularly for<br />

species with short-lived seed banks (Pywell et al. 2002). Therefore studies on seed dispersal<br />

have become an important issue in plant ecology in general <strong>and</strong> restoration management in<br />

particular. In the framework of nature conservation <strong>and</strong> restoration, seed dispersal by livestock<br />

has been examined by Bakker et al. (2005), Malo <strong>and</strong> Suarez (1995a), Mouissie et al. (2005),<br />

Couvreur et al. (2005), Cosyns et al. (2005b) who showed the potential of endozoochorous seed<br />

dispersal in a semi-natural l<strong>and</strong>scape. The contribution of large herbivores to ecological<br />

restoration sites can be determined by gathering quantitative information on the dung seed<br />

content. Predicting seed input by large herbivores through germination test of seeds in dung is<br />

needed to figure out the role of large herbivores.<br />

Seed dispersal contributes to the diversity of plant species in the seed bank. Large herbivores<br />

consume large numbers of seed (Malo <strong>and</strong> Suarez 1995b) <strong>and</strong> move long distances, due to their<br />

extensive home-ranges (Haskell et al. 2002). Bakker <strong>and</strong> Olff (2003) found 183.1 ± 21.1<br />

seedlings per 100 gram dung of cattle in September <strong>and</strong> 9.8 ± 2.1 seedlings per 100 gram dung<br />

of rabbit in the same month. Most nature conservation areas in The Netherl<strong>and</strong>s consist of open<br />

area (heathl<strong>and</strong> <strong>and</strong> grassl<strong>and</strong>) <strong>and</strong> woodl<strong>and</strong>. In some part of this area, large herbivores like<br />

cattle <strong>and</strong> horses are used as management tools. Defoliation, treading <strong>and</strong> excretion are direct<br />

effects of large herbivores on the vegetation. Modification of the plant environment is indeed an<br />

indirect effect (Bokdam <strong>and</strong> Gleichman 2000). Usually large herbivores graze in open areas but<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.W Saragih et al. 2010. Effects of endozoochorous seed dispersal on the soil seed bank <strong>and</strong> vegetation in the woodl<strong>and</strong> area<br />

160<br />

they ruminate <strong>and</strong> defecate in woodl<strong>and</strong> area. In the Wolfhezerheide cattle spend most nonforaging<br />

time (56%) in woodl<strong>and</strong> area (Bokdam 2003). This causes seed disperse from open<br />

areas that are dominated by ‘lawn species’ to woodl<strong>and</strong> areas that are dominated by forest<br />

species. The large herbivores may have an important influence on the soil seed bank in the<br />

woodl<strong>and</strong> area through endozoochory. Pakeman et al. (2002) gave evidence of a relation<br />

between soil seed bank <strong>and</strong> dung seed content. Dung deposition by cattle influences plant<br />

dynamics through the combined effects of seed input <strong>and</strong> gap formation i.e. dung itself as a<br />

potential regeneration site or as a precursor of gaps by suppressing vegetation (Cosyns et al.<br />

2005a). The present of grassl<strong>and</strong> species in the gaps of woodl<strong>and</strong> area confirmed the<br />

contribution of endozoochory to the soil seed bank. In the Wolfhezerheide Nature Reserve,<br />

cattle <strong>and</strong> rabbits are the herbivores that play an important role in seed dispersal. Large numbers<br />

of seeds are potentially dispersed via cattle <strong>and</strong> rabbit dung. In this study, endozoochorous seed<br />

dispersal affect species richness in the vegetation directly <strong>and</strong> soil seed bank indirectly.<br />

Different seasons contribute different species richness <strong>and</strong> the number of seed in the soil seed<br />

bank.<br />

2. Methodology<br />

2.1 Study site <strong>and</strong> experimental design<br />

This study was carried out in the nature reserve ’The Wolfhezerheide’ (120 ha)<br />

(Naturmonumenten 1996; (Bokdam 2003). This area is located in the Veluwe in the centre of<br />

The Netherl<strong>and</strong>s (51º 47’N; 5º 41’E) <strong>and</strong> owned by Vereniging Natuurmonumenten. The<br />

reserve consists of 30 ha forest <strong>and</strong> 90 ha open area. The soils are mainly acid (pH KCl 3–3.5)<br />

<strong>and</strong> podzolic, developed in Pleistocene fluvio-glacial s<strong>and</strong> <strong>and</strong> cover s<strong>and</strong> (Bokdam 2003).<br />

Calluna vulgaris, Deschampsia flexuosa <strong>and</strong> Molinia caerulea are dominant plant species in the<br />

open area. Pinus sylvestris <strong>and</strong> Betula spp. dominate the forest area. Since 1983, grazing<br />

management has been applied in this area. The general approach in this research was<br />

observational by using a fence line study technique where the fence is used to divide the study<br />

area in a grazed <strong>and</strong> a non-grazed area.<br />

2.2 Data collections<br />

Data collection of this study consists of vegetation composition, cattle dung <strong>and</strong> soil seed bank.<br />

Vegetation data were collected from grazed <strong>and</strong> ungrazed woodl<strong>and</strong> sites. Pairs of sites with a<br />

size of 5 x 5 m (25 m 2 ) were chosen opposite to each other on either side of the fence, with ten<br />

replicates resulting in ten relevés in the grazed area <strong>and</strong> ten relevés in the ungrazed area. The<br />

present of each species in a relevés was estimated by its cover <strong>and</strong> used to assess the similarity<br />

between vegetation in the grazed <strong>and</strong> ungrazed woodl<strong>and</strong> area. The numbers of species in both<br />

sites are also compared to find out the differences in plant species diversity. Cattle dung<br />

samples were collected from two seasons. Dung sample (141 gram) was air dried in the<br />

greenhouse at the prevailing temperature (25-28 o C) for 2 weeks before storing them in the<br />

cooler at 5 o C for two weeks. After the cool period, trays were placed in the greenhouse for the<br />

germination test period. In addition, the soil seed bank was sampled with composite samples of<br />

the litter layer plus the top five centimeter of the mineral soil were collected from the plot where<br />

the vegetation was studied. Each composite sample consists of ten cores (2.5 cm in diameter x<br />

25 cm deep) is taken r<strong>and</strong>omly within a plot. Soil samples were sieved to remove coarse gravel,<br />

roots <strong>and</strong> vegetation. The samples were spread in trays with one cm thickness (672 g or 0.8 L)<br />

on top of a three cm thick layer of coarse river s<strong>and</strong>. A cold treatment (6 o C) was given to the<br />

samples for two weeks to break dormancy <strong>and</strong> improved germination of the seeds (Olff et al.<br />

1994).<br />

2.3 Seedling emergence technique<br />

For germination test in the green house, the weight <strong>and</strong> volume of the soil <strong>and</strong> the dry dung<br />

samples were determined. After cold stratification treatment, the trays were placed in a<br />

greenhouse with day temperature of 20 o C <strong>and</strong> night temperature of 18 o C (12/12 hr day-night<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.W Saragih et al. 2010. Effects of endozoochorous seed dispersal on the soil seed bank <strong>and</strong> vegetation in the woodl<strong>and</strong> area<br />

161<br />

light regime) <strong>and</strong> air humidity 70%. The trays were placed under a hood of plastic foil to reduce<br />

evaporation <strong>and</strong> contamination of seed within the greenhouse. The emerging seedlings were<br />

identified by using identification keys (Muller 1978).<br />

2.4 Data Analysis<br />

The data were analyzed using SPSS 12.01, a statistical software for Windows (SPSS 2003).<br />

3. Results<br />

3.1 Cover, species richness, species composition<strong>and</strong> lawn species in the vegetation<br />

The average cover of vegetation in the grazed area is higher than in the ungrazed area (48.865%<br />

vs. 69.972 %). A paired sample T-test executed to figure out different on cover vegetation<br />

between them. The result showed significant difference on cover vegetation between vegetation<br />

ungrazed <strong>and</strong> grazed woodl<strong>and</strong> area (t = -2.481; N = 10; P = 0.035). The species <strong>and</strong> total<br />

number of species in the pooled of vegetation is higher in the grazed area than in the ungrazed<br />

area (32 vs. 20 species). The Pair T-test tested the significance of the difference between the<br />

average species richness of ungrazed <strong>and</strong> grazed plots. It showed a significant difference in the<br />

number species between two treatments (areas) (t = -4.017; N = 10; P = 0.003). The average<br />

number of species per plot in the grazed area (11.3 ± 1.585) is higher than in the ungrazed area<br />

(5.5 ± 0.373). Ungrazed <strong>and</strong> grazed areas had 35 species in total <strong>and</strong> 17 species in common.<br />

Lawn species in this study are herbs <strong>and</strong> dwarf shrubs with Ellenberg value for light larger than<br />

six. In total, 14 lawn species found in both ungrazed <strong>and</strong> grazed vegetation of woodl<strong>and</strong> area.<br />

Both of them shared five of lawn species (Agrostis capillaries, Deschampsia flexuosa, Galium<br />

saxatile, Rubus fruticosus agg, Taraxacum officinale). Nine species out of 14 were found only<br />

in the grazed area <strong>and</strong> two species (Holcus lanatus <strong>and</strong> Senecio jacobea) were dispersed by<br />

endozoochory. 31.47% is total cover of lawn species in ungrazed area <strong>and</strong> 65.08% in grazed<br />

area. Total frequency of lawn species in ungrazed is 2.1 <strong>and</strong> 5.2 in grazed area.<br />

3.2 Seed density, species richness, species composition <strong>and</strong> lawn species of summer<br />

<strong>and</strong> winter dung<br />

A total of 986 seedlings emerged from the winter dung, while 7680 seedlings sprouted from the<br />

summer dung samples (six litres, dry weight 2115 gr). The parametric test for two independent<br />

samples showed significant difference on the number of seeds between winter <strong>and</strong> summer (t=-<br />

5.504; N=20; P=0.000). Total number of species of winter dung samples is higher than in<br />

summer dung samples (25 vs. 17 species). A parametric test for two independent samples T-test<br />

showed a significant difference in the average number of species between winter <strong>and</strong> summer<br />

dung (F = 0.005; N = 25; P = 0.009). In the species richness, in total 34 species were found in<br />

winter <strong>and</strong> summer dung one <strong>and</strong> only eight species in common. From a total of 25 species in<br />

the winter dung, 17 species did not occur in the summer dung. Nine species of summer dung did<br />

not occur in the winter dung. In total, 24 lawn species were found in both seasons. Six lawn<br />

species were shared in ungrazed <strong>and</strong> grazed (Calluna vulgaris, Holcus lanatus, Lolium perenne,<br />

Polygonum aviculare, Senecio jacobae, <strong>and</strong> Veronica arvensis). 13 species (out of 24) were<br />

exclusively found in the winter dung <strong>and</strong> five exclusively in the summer dung. The number of<br />

seeds of lawn species was higher (509.73) in the summer dung than in the winter (92.4), but the<br />

frequency was higher in winter dung (8.1) than in the summer (5).<br />

3.3 Seed density, species richness,species composition <strong>and</strong> lawn in the soil seed<br />

banks.<br />

A total of 601 seedlings (representing 22 species) emerged from soil samples ungrazed forest<br />

area, 360 seedlings with 23 species from the grazed forest area. Six species (Genista anglica,<br />

Jasione montana, Lamium purpureum, Plantago major <strong>and</strong> Sonchus oleraceus <strong>and</strong> Polygonum<br />

persicaria) were only found in the grazed woodl<strong>and</strong>. Five species (Hypericum perforatum,<br />

Oxalis corniculata, Rumex acetosella, Vaccinuium mytillus <strong>and</strong> Veronica arvensis) occurred<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.W Saragih et al. 2010. Effects of endozoochorous seed dispersal on the soil seed bank <strong>and</strong> vegetation in the woodl<strong>and</strong> area<br />

162<br />

only in the ungrazed woodl<strong>and</strong>. In total, 19 lawn species found in both soil seed bank. Two<br />

lawn species (Cirsium vulgare <strong>and</strong> Veronica arvensis) only present in ungrazed area while five<br />

species (Genista anglica, Juncus buffonius agg, Plantago major, Sonchus oleraceus <strong>and</strong><br />

Lamium purpureum) only present in grazed area. However only Jasione montana, Lamium<br />

purpureum & Sonchus oleraceus were found in the dung <strong>and</strong> those species dispersed during the<br />

winter. Number of seeds is higher in the soil seed bank ungrazed area than soil seed bank grazed<br />

area but frequency of lawn species in soil seed bank of grazed area is higher than ungrazed area.<br />

Total frequency lawn species in grazed area (4.6) is higher than ungrazed area (4.2)<br />

4. Discussion<br />

4.1 The effects of cattle on woodl<strong>and</strong> vegetation<br />

Cattle play an important role in the vegetation which show by species richness as well as cover<br />

was higher in the grazed woodl<strong>and</strong> vegetation than ungrazed ones. Many of these are the<br />

characteristic species of pasture <strong>and</strong> heathl<strong>and</strong> (Senecio jacobaea, Cerastium fontanum,<br />

Danthonia decumbens, Plantago lanceolata, Poa pratensis, Ranunculus repens, Rumex<br />

acetosella, Trifolium repens), clearings (Senecio sylvaticus) or fringes (Rubus idaeus, Teucrium<br />

scorodonia). Grazing reduced the cover of grazing-sensitive species: Ceratocapnos claviculata,<br />

Dryopteris carthusiana, <strong>and</strong> Vaccinium myrtillus. Agrostis capillaris has the same average<br />

cover in grazed <strong>and</strong> ungrazed but species frequency in grazed area is higher than in ungrazed<br />

area. In addition, Deschampsia flexuosa <strong>and</strong> Galium saxatile had the same species frequencies<br />

but average cover of species was higher in grazed vegetation. This evidenced present<br />

contribution of cattle-dispersed lawn species in the woodl<strong>and</strong> area by creating understory gaps,<br />

stimulate germination <strong>and</strong> growth of lawn species in the understory vegetation of the woodl<strong>and</strong>.<br />

Seven species of dung samples (Erigeron Canadensis, Epilobium sp, Juncus bufonius agg,<br />

Lolium perenne, Medicago lupulina, Ranunculus acris) were absent from aboveground<br />

vegetation in the grazed area due unable to cope with the environmental condition <strong>and</strong><br />

predators. It seemed that the number of species aboveground is not influenced by the number of<br />

species in the soil seed bank, as the numbers of species in the soil seed bank depend on the<br />

number of persistent seeds. The floristic composition of the seed bank is not a mirror reflecting<br />

the vegetation composition, but rather a record of a long-term turnover of species <strong>and</strong> many<br />

different events which in various periods of time influence the input <strong>and</strong> output of seeds<br />

(Falinska 1998).<br />

4.2 Contribution of endozoochorous seed dispersal to lawn species establishment<br />

The dung deposited by herbivores (cattle) increased the diversity in the grazed woodl<strong>and</strong><br />

vegetation that could cause by cattle facilitated the arrival of species from open area to the<br />

woodl<strong>and</strong> area <strong>and</strong> good condition provided by dung for germination. Dung creates places free<br />

from vegetation (gaps) with high nutrient availability (Dai 2000) <strong>and</strong> dung with sufficient<br />

water- retention capacity provide a safe site for a number of species particularly those from<br />

nutrient-rich habitat that able to grow fast <strong>and</strong> root in underlying soil (Mouissie 2005). The right<br />

conditions for germination <strong>and</strong> establishment, which could be linked to the change in nutrient<br />

concentrations, were provided by dung <strong>and</strong> might have also generated positive effects on the<br />

germination of some of the species present in the soil seed bank (Traba et al. 2003). Eleven<br />

light dem<strong>and</strong>ing species (Danthonia decumbens, Poa pratensis, Rumex acetocella, Ranunculus<br />

repens, Rosa canina Rubus idaeus, Sorbus ocuparia, Sinecio sylvaticus, Sambucus nigra,<br />

Teucrium scorodonia, Trifolium repens) from grazed vegetation were absent in the species dung<br />

sample. Those species might be unpalatable or not providing seeds during the study period.<br />

Endozoochorous seed dispersal explained the exclusive occurrence of Cerastium fontanum,<br />

Holcus lanatus, Plantago lanceolata, Poa trivialis <strong>and</strong> Senecio jacobaea the in grazed<br />

vegetation. Most of them were exclusively present in the summer dung, except for Holcus<br />

lanatus which was found in both seasons. Those species are light dem<strong>and</strong>ing (L- value≥6). Total<br />

cover <strong>and</strong> total frequency of lawn species is higher in vegetation of grazed area than in the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.W Saragih et al. 2010. Effects of endozoochorous seed dispersal on the soil seed bank <strong>and</strong> vegetation in the woodl<strong>and</strong> area<br />

163<br />

ungrazed area. This might indicate contribution of large herbivore on the establishment of lawn<br />

species.<br />

4.3 The contribution of endozoochous seed dispersal to the soil seed bank<br />

The seed density in ungrazed woodl<strong>and</strong> are is higher than in grazed woodl<strong>and</strong>. The expectation<br />

was in contrast because cattle contribute the number of seeds into the grazed area. One reason<br />

for the higher number of seed in ungrazed soil could be the thickness of litter layer. The litter<br />

layer in the ungrazed (36 cm) area is thicker than in the grazed area (30.5 cm). Litter increased<br />

seed longevity (Rotundo <strong>and</strong> Aguiar 2005) <strong>and</strong> when the litter layer is thick, this can lead to the<br />

conservation of seed in the soil (Jensen 1998). Furthermore, trampling by cattle can bury seeds<br />

deeply so they were not present in the top soil samples (30 cm deep). McDonald et al. 1996<br />

stated that grazing allows the incorporation of seeds deeply in the soil. Twelve species of<br />

summer dung sample absent from soil seed bank ungrazed <strong>and</strong> grazed woodl<strong>and</strong> area. The<br />

summer dung species needed more light (L Ellenberg’s value >6) <strong>and</strong> have the same<br />

requirement levels of moisture <strong>and</strong> nitrogen. Winter dung might contribute more to the species<br />

richness of the soil seed bank than the summer dung did. This could be the case as during the<br />

winter, limited foods were available for the cattle which led them to consume more different<br />

plants species. On the contrary in the summer, the cattle preferred the most palatable<br />

species, mostly grasses which were abundant during this season. Grass species have a<br />

low seed longevity index <strong>and</strong> low densities in the soil. Total seed density of lawn<br />

species was higher in soil seed bank ungrazed area than grazed area, however total<br />

frequency of lawn species in soil seed bank grazed area is higher than ungrazed. This<br />

might indicate role of cattle on distribution of lawn species which showed by more lawn<br />

species was found in soil seed bank grazed area than ungrazed area. Jasione montana,<br />

Lamium purpureum & Sonchus oleraceus dispersed by cattle during the winter <strong>and</strong> only<br />

found in soil seed bank grazed area.<br />

References<br />

Bakker, C., H. F. d. Graaf, W. H. O. Ernst, <strong>and</strong> P. M. v. Bodegom. 2005. Does the seed bank<br />

contribute to the restoration of species-rich vegetation in wet dune slacks Applied<br />

Vegetation Science. 2005; 8:39-48.<br />

Bakker, E. E. S., <strong>and</strong> H. H. Olff. 2003. Impact of different-sized herbivores on recruitment<br />

opportunities for subordinate herbs in grassl<strong>and</strong>s. Journal Of Vegetation Science 14:465.<br />

Bokdam, J., <strong>and</strong> J. M. Gleichman. 2000. Effects of grazing by free-ranging cattle on vegetation<br />

dynamics in a continental north-west European heathl<strong>and</strong>. Journal Of Applied Ecology<br />

37:415-431.<br />

Bokdam, J. S. 2003. Nature conservation <strong>and</strong> grazing management: free-ranging cattle as a<br />

driving force for cyclic vegetation succession, Wageningen. The Netherl<strong>and</strong>s.<br />

Cosyns, E., A. Delporte, L. Lens, <strong>and</strong> M. Hoffmann. 2005a. Germination success of temperate<br />

grassl<strong>and</strong> species after passage through ungulate <strong>and</strong> rabbit guts. Journal Of Ecology<br />

93:353-361.<br />

Cosyns, E. E., S. S. Claerbout, I. I. Lamoot, <strong>and</strong> M. M. Hoffmann. 2005b. Endozoochorous seed<br />

dispersal by cattle <strong>and</strong> horse in a spatially heterogeneous l<strong>and</strong>scape. Plant ecology<br />

178:149.<br />

Couvreur, M., E. Cosyns, M. Hermy, <strong>and</strong> M. Hoffmann. 2005. Complementarity of epi- <strong>and</strong><br />

endozoochory of plant seeds by free ranging donkeys. Ecography 28:37-48.<br />

Dai, X. 2000. Impact of cattle dung deposition on the distribution pattern of plant species in an<br />

alvar limestone grassl<strong>and</strong>. Journal Of Vegetation Science 11:715.<br />

Falinska, K. 1998. Long-term changes in size <strong>and</strong> composition of seed bank during succession:<br />

From meadow to forest. Acta Societatis Botanicorum Poloniae 67:301.<br />

Haskell, J. P., M. E. Ritchie, <strong>and</strong> H. Olff. 2002. Fractal geometry predicts varying body size<br />

scaling relationships for mammal <strong>and</strong> bird home ranges. Nature 418:527-530.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.W Saragih et al. 2010. Effects of endozoochorous seed dispersal on the soil seed bank <strong>and</strong> vegetation in the woodl<strong>and</strong> area<br />

164<br />

Jensen, K. 1998. Species composition of soil seed bank <strong>and</strong> seed rain of ab<strong>and</strong>oned wet<br />

meadows <strong>and</strong> their relation to aboveground vegetation. Flora 193:345-359.<br />

Malo, J. E. 2000. Hardseededness <strong>and</strong> the accurancy of seed bank estimates obtained through<br />

germination. web. ecology 1:70-75.<br />

Malo, J. E., <strong>and</strong> F. Suarez. 1995a. Establishment Of Pasture Species On Cattle Dung - The Role<br />

Of Endozoochorous Seeds. Journal Of Vegetation Science 6:169-174.<br />

Malo, J. E., <strong>and</strong> F. Suarez. 1995b. Herbivorous Mammals As Seed Dispersers In A<br />

Mediterranean Dehesa. Oecologia 104:246-255.<br />

McDonald, A. W., J. P. Bakker, <strong>and</strong> K. Vegelin. 1996. Seed bank classification <strong>and</strong> its<br />

importance for the restoration of species-rich flood-meadows. Journal Of Vegetation<br />

Science 7:157-164.<br />

Mouissie, A. M., P. Vos, H. M. C. Verhagen, <strong>and</strong> J. P. Bakker. 2005. Endozoochory by freeranging,<br />

large herbivores: ecological correlates <strong>and</strong> perspectives for restoration. Basic<br />

<strong>and</strong> Applied Ecology. 2005; 6:547-558.<br />

Muller, F. M. 1978. Seedlings of the North-Western European lowl<strong>and</strong>: a flora of seedlings,<br />

Wageningen.<br />

Olff, H., D. M. Pegtel, J. M. Vangroenendael, <strong>and</strong> J. P. Bakker. 1994. Germination Strategies<br />

During Grassl<strong>and</strong> Succession. Journal Of Ecology 82:69-77.<br />

Pakeman, R. J., G. Digneffe, <strong>and</strong> J. L. Small. 2002. Ecological correlates of endozoochory by<br />

herbivores. Functional Ecology 16:296-304.<br />

Pywell, R. R. F., J. J. M. Bullock, A. A. Hopkins, K. K. J. Walker, T. T. H. Sparks, M. M. J. W.<br />

Burke, <strong>and</strong> S. S. Peel. 2002. Restoration of species-rich grassl<strong>and</strong> on arable l<strong>and</strong>:<br />

assessing the limiting processes using a multi-site experiment. The Journal of applied<br />

ecology 39:294.<br />

Rotundo, J. L., <strong>and</strong> M. R. Aguiar. 2005. Litter effects on plant regeneration in arid l<strong>and</strong>s: a<br />

complex balance between seed retention, seed longevity <strong>and</strong> soil-seed contact. Journal<br />

Of Ecology 93:829-838.<br />

SPSS, I. 2003. SPSS 12.01 for windows.<br />

Traba, J., C. Levassor, <strong>and</strong> B. Peco. 2003. Restoration of Species Richness in Ab<strong>and</strong>oned<br />

Mediterranean Grassl<strong>and</strong>s: Seeds in Cattle Dung. Restoration Ecology 11:378-384.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.R.P. Williamson et al. 2010. Anthropogenic l<strong>and</strong>scape changes <strong>and</strong> the conservation of woodl<strong>and</strong> caribou<br />

165<br />

Anthropogenic l<strong>and</strong>scape changes <strong>and</strong> the conservation of woodl<strong>and</strong><br />

caribou in British Columbia, Canada<br />

Libby R.P. Williamson * , Chris J. Johnson 1 & Dale R. Seip 2<br />

1 Ecosystem Science <strong>and</strong> Management Program, University of Northern British Columbia,<br />

3333 University Way, Prince George, British Columbia, V2N 4Z9, Canada.<br />

2 British Columbia Ministry of <strong>Forest</strong>s, 1011 4 th Ave, Prince George, British Columbia,<br />

V2L 3H9, Canada.<br />

Abstract<br />

Conservation <strong>and</strong> management of caribou across Canada is now a high priority due to increasing<br />

rates of decline associated with greater levels of l<strong>and</strong>scape change <strong>and</strong> anthropogenic disturbance.<br />

Two of the most pressing threats to caribou are habitat loss <strong>and</strong> predation; both effects are a direct<br />

result of large-scale forestry, energy, <strong>and</strong> mineral development. Mountain, northern <strong>and</strong> boreal<br />

caribou are three ecotypes of woodl<strong>and</strong> caribou (Rangifer tar<strong>and</strong>us caribou) found in British<br />

Columbia. Across most of the province, the federal government has listed all three ecotypes of<br />

woodl<strong>and</strong> caribou as “threatened”, but recovery planning for caribou in the province has met with<br />

only limited success. This literature review will provide a foundation to further underst<strong>and</strong> how<br />

l<strong>and</strong>scape change is contributing to declines in woodl<strong>and</strong> caribou populations <strong>and</strong> how continued<br />

research will aid in favorable decisions for the long-term survival <strong>and</strong> management of this keystone<br />

species of the North.<br />

Keywords: Rangifer tar<strong>and</strong>us caribou, industrial disturbance, population decline, conservation,<br />

predation<br />

1. Introduction<br />

The North would not be what it is today, without caribou. Rangifer tar<strong>and</strong>us is considered to be a<br />

single species throughout the world <strong>and</strong> includes both domestic reindeer <strong>and</strong> wild caribou (Hummel<br />

<strong>and</strong> Ray 2008). The term “reindeer” often refers to domesticated members in Europe, while<br />

“caribou” is reserved for wild members of the same species in North America. Caribou <strong>and</strong><br />

reindeer reside in most of the world’s north, above the 50 th parallel in North America <strong>and</strong> the 62 nd<br />

parallel in Eurasia, respectively (Hummel <strong>and</strong> Ray 2008).<br />

In the northern extent of their range, caribou congregate in large social communities, take<br />

part in major migrations from their wintering to calving grounds, <strong>and</strong> are commonly referred to as<br />

“barren-ground” caribou. Towards the south, characteristics change <strong>and</strong> herds remain at low<br />

densities, have shortened migrations, <strong>and</strong> are often classified as “sedentary”. Caribou have been<br />

classified into three primary ecotypes across North America based upon their behavioral <strong>and</strong><br />

ecological differences: mountain, boreal forest, <strong>and</strong> migratory tundra (Hummel <strong>and</strong> Ray 2008).<br />

Woodl<strong>and</strong> caribou (Rangifer tar<strong>and</strong>us caribou) is a sub-species found across each of the three<br />

ecotypes of North American caribou.<br />

* Corresponding author. Tel.:1(250)960 6739<br />

Email address: williame@unbc.ca<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


L.R.P. Williamson et al. 2010. Anthropogenic l<strong>and</strong>scape changes <strong>and</strong> the conservation of woodl<strong>and</strong> caribou<br />

166<br />

2. General Ecology of Woodl<strong>and</strong> Caribou in British Columbia<br />

Similar to the federal system for North American caribou, British Columbia (B.C.) classifies<br />

woodl<strong>and</strong> caribou into three ecotypes: mountain, northern <strong>and</strong> boreal (Heard <strong>and</strong> Vagt 1998).<br />

Mountain caribou rely on old-growth subalpine <strong>and</strong> rugged alpine habitats in the central <strong>and</strong><br />

southeastern portions of the province. During winter, these caribou forage on abundant arboreal<br />

lichens (Bryoria spp. <strong>and</strong> Alectoria sarmentosa) as deep snow restricts there access to terrestrial<br />

lichens or vascular plants (Stevenson <strong>and</strong> Hatler 1985; Seip <strong>and</strong> Cichowski 1996; Jones et al. 2007).<br />

Moving to higher elevations to access forage in the winter is an effective strategy for avoiding<br />

predators (Seip <strong>and</strong> McLellan 2008).<br />

Caribou of the northern ecotype prefer terrestrial lichens (Cladina mitis <strong>and</strong> Cladonia spp.)<br />

that are found in lower-elevation pine forests or alpine habitats (Heard <strong>and</strong> Vagt 1998; Johnson et<br />

al. 2004; Jones et al. 2007). Depending on snow conditions <strong>and</strong> lichen abundance, these caribou<br />

will also forage on arboreal lichens (Bryoria spp.) during the winter months (Johnson et al. 2004).<br />

Northern caribou have highly variable wintering strategies between years, populations <strong>and</strong><br />

individuals; some caribou will winter on high, wind-swept alpine ridges, while others prefer<br />

wintering in lower-elevation pine-lichen forests (Bergurud 1978; Terry <strong>and</strong> Wood 1999; Johnson et<br />

al. 2004; Jones et al. 2007).<br />

The boreal ecotype is found in the northeastern portion of the province <strong>and</strong> prefers black<br />

spruce (Picea mariana) fen/bog complexes, <strong>and</strong> tends to avoid well-drained areas (Stuart-Smith et<br />

al. 1997; Dzus 2001). A lack of topographic relief prevents boreal caribou from making<br />

elevational migrations as demonstrated by the mountain <strong>and</strong> northern ecotypes (Stuart-<br />

Smith et al. 1997; Culling et al. 2006). Ground lichens (Cladina stellaric, C. mitis <strong>and</strong> C.<br />

rangiferina) become the dominant food source in winter (Schaefer 2008). Boreal caribou now<br />

occupy less than half of their historical range across the continent (Schaefer 2008).<br />

3. Threats to Populations of Woodl<strong>and</strong> Caribou<br />

The abundance <strong>and</strong> distribution of woodl<strong>and</strong> caribou has been on the decline across North America<br />

since European advancement <strong>and</strong> colonization (Bergerud 1974; Seip 1992; Vors et al. 2007).<br />

Conservation <strong>and</strong> management of caribou across Canada is now a high priority due to increasing<br />

rates of decline associated with greater levels of l<strong>and</strong>scape change <strong>and</strong> anthropogenic disturbance<br />

(Wittmer 2004; Vors <strong>and</strong> Boyce 2009). Two of the most pressing threats to caribou are habitat loss<br />

<strong>and</strong> predation; both effects are a direct result of large-scale forestry, energy, <strong>and</strong> mineral<br />

development (Schaefer 2003; Vors et al. 2007).<br />

Predation is suggested as the leading cause of mortality for woodl<strong>and</strong> caribou in North<br />

America with wolves serving as the primary predator in this multi-carnivore ecosystem (Bergerud<br />

1974; Fuller <strong>and</strong> Keith 1981; Stewart-Smith et al. 1997; Kinley <strong>and</strong> Apps 2001; Gustine et al.<br />

2006). Caribou are thought to minimize the risk of predation by using habitats that are spatially<br />

separated from predators (Bergerud et al. 1984; Stuart-Smith et al. 1997; Cumming et al. 1996;<br />

Johnson et al. 2004; Latham 2009). During the calving season, for example, caribou in the<br />

mountainous regions of B.C. <strong>and</strong> west-central Alberta reduce predation by moving to higher<br />

elevations (Bergurud et al. 1984; Seip 1992; Johnson et al. 2002). Similarly, boreal caribou in<br />

northeastern B.C. use predator refugia such as lakeshores, small isl<strong>and</strong>s of mature black spruce,<br />

thick alder st<strong>and</strong>s saturated with water, <strong>and</strong> old burn sites adjacent to wetl<strong>and</strong>s during the calving<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


L.R.P. Williamson et al. 2010. Anthropogenic l<strong>and</strong>scape changes <strong>and</strong> the conservation of woodl<strong>and</strong> caribou<br />

167<br />

season (Culling et al. 2006). Studies in Ontario also show an increased use of isl<strong>and</strong>s <strong>and</strong> lake<br />

shorelines during the spring (Bergerud 1985; Cumming <strong>and</strong> Beange 1987).<br />

Many declines in ungulate populations have been attributed to habitat degradation (Leopold<br />

<strong>and</strong> Darling 1953; Bradshaw <strong>and</strong> Hebert 1996). L<strong>and</strong>scape change <strong>and</strong> an increase in the<br />

abundance of other ungulate species now limit the ability of caribou to effectively space-away from<br />

predators (Rempel et al. 1997; Wittmer 2004; Latham 2009). Since the early 1900s, moose (Alces<br />

alces) have exp<strong>and</strong>ed their distribution throughout B.C. resulting in a numerical <strong>and</strong> distributional<br />

response of wolves (Bergerud <strong>and</strong> Elliot 1986; Seip 1992; Spalding 1990). Known as “apparent<br />

competition”, primary prey species such as deer <strong>and</strong> moose do not compete directly with caribou for<br />

forage or space, but support larger number of wolves that prey on caribou opportunistically (Holt<br />

1977; Bergerud <strong>and</strong> Elliot 1986; Seip 1992; Wittmer et al. 2005; Latham 2009).<br />

Recent studies in B.C. <strong>and</strong> Alberta have demonstrated that industrial activities <strong>and</strong><br />

associated linear features, including roads, trails, geophysical exploration lines, pipelines, electrical<br />

right-of-ways, cutblocks, <strong>and</strong> oil <strong>and</strong> gas wells can negatively affect woodl<strong>and</strong> caribou populations<br />

(Bradshaw et al. 1997, James <strong>and</strong> Stuart-Smith 2000, Smith et al. 2000, Dyer et al. 2001, Sorensen<br />

et al. 2008). These features can alter the movements, distributions, <strong>and</strong> population dynamics of<br />

both caribou <strong>and</strong> wolves. Timber harvesting is one of the primary agents of habitat change. Largescale<br />

harvesting reduces the amount of habitat for caribou <strong>and</strong> increases the area of early<br />

successional forests favored by moose <strong>and</strong> other ungulate species (Fuller <strong>and</strong> Keith 1981, Rempel et<br />

al. 1997, Johnson et al 2004, Nitschke 2008). There is rising concern about how advancing threats<br />

from industry are influencing the ability of sub-populations to survive. Small populations, like the<br />

mountain caribou in the southern portions of B.C., have become isolated from neighboring herds<br />

<strong>and</strong> are at greater risk of extirpation from r<strong>and</strong>om variation <strong>and</strong> stochastic events (Heard <strong>and</strong> Vagt,<br />

1998).<br />

Recent evidence suggests that disturbance from recreational activities may also be<br />

contributing to the decline of northern <strong>and</strong> mountain caribou herds (Seip et al. 2007). Recreational<br />

tenures for commercial purposes have increased in both area <strong>and</strong> number across the province; these<br />

tenures allow greater access into caribou habitat for snowmobiles <strong>and</strong> helicopter ski (heli-ski)<br />

operations (McNay <strong>and</strong> Giguere 2008). As recreational activities exp<strong>and</strong> across alpine areas used<br />

by caribou, herds are forced into less favorable habitats that increase accidental mortalities from<br />

avalanches, increase energy dem<strong>and</strong>s used to move across steep terrain <strong>and</strong> deep snow, as well as<br />

increase the risk of predation (Seip et al. 2007).<br />

4. Management of Woodl<strong>and</strong> Caribou<br />

Across most of B.C., the federal government has listed all three ecotypes of woodl<strong>and</strong> caribou as<br />

“threatened”. Management goals continue to focus on monitoring population levels, restoring <strong>and</strong><br />

maintaining appropriate sex <strong>and</strong> age ratios, <strong>and</strong> include a required inspection for all human<br />

harvested caribou (populations of northern <strong>and</strong> boreal ecotypes that are not considered threatened).<br />

In order to continue developing the most effective conservation strategies for species-at-risk,<br />

professionals must rely on the philosophy of their discipline, scientific <strong>and</strong> traditional ecological<br />

knowledge (TEK) from indigenous communities, as well as educated opinions that come from<br />

expert peers (Stephenson 1982; Mountain Caribou Technical Advisory Committee 2002).<br />

All provinces <strong>and</strong> territories in Canada must adhere to federal endangered species<br />

legislation. For species listed as “endangered” or “threatened”, SARA (Species at Risk Act 2003)<br />

requires that the responsible jurisdiction, or authority, develop <strong>and</strong> implement a Recovery Action<br />

Plan. These recovery plans address immediate threats to the species <strong>and</strong> protects or enhances the<br />

species residence <strong>and</strong> critical habitat. Two advisory committees were formed in B.C. to provide<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


L.R.P. Williamson et al. 2010. Anthropogenic l<strong>and</strong>scape changes <strong>and</strong> the conservation of woodl<strong>and</strong> caribou<br />

168<br />

strategic direction for the recovery of caribou in the province <strong>and</strong> multiple Recovery<br />

Implementation Groups (RIGs) develop Recovery Action Plans that address the specific<br />

conservation needs of collections of herds.<br />

Recovery planning for caribou in B.C. has met with only limited success. The Mountain<br />

Caribou Technical Advisory Committee has produced a document outlining the threats <strong>and</strong> a<br />

strategy for the recovery of mountain caribou in B.C (Mountain Caribou Technical Advisory<br />

Committee 2002). Both the North-Central <strong>and</strong> the Hart <strong>and</strong> Cariboo Mountains RIGs have<br />

developed recovery action plans that make specific conservation recommendations for establishing<br />

self-sustaining populations of caribou. In 2003, the Central Rocky Mountain RIG was temporarily<br />

suspended <strong>and</strong> there are no committees designated for implementing recovery strategies for boreal<br />

caribou in B.C.<br />

5. Future Directions in Research<br />

As we progress into the future, continued research in caribou ecology remains crucial. Populations<br />

of woodl<strong>and</strong> caribou should be continuously monitored <strong>and</strong> surveyed across Canada. If we are<br />

unclear on the health <strong>and</strong> overall status of caribou, it challenges our underst<strong>and</strong>ing of how to<br />

properly implement habitat protection. Studies should continue looking into relationships between<br />

l<strong>and</strong>scape disturbance <strong>and</strong> habitat change, woodl<strong>and</strong> caribou <strong>and</strong> increasing numbers of elk, deer,<br />

<strong>and</strong> moose populations. Additional studies focusing on the predator-prey dynamics of wolves with<br />

their primary prey will provide insight into alternate management strategies where caribou<br />

populations are dramatically declining. Moose-wolf interactions, for example, are a key component<br />

that shape caribou population dynamics. Future research should also include predation studies that<br />

specifically focus on the spatial distribution, habitat selection <strong>and</strong> hunting patterns of wolves <strong>and</strong><br />

other influential carnivores (e.g., bears <strong>and</strong> cougars) living in caribou habitat. Intensive efforts in<br />

the conservation of woodl<strong>and</strong> caribou will continue to grow as changing predator-prey relations are<br />

exacerbated by l<strong>and</strong>scape change due to advancing human developments.<br />

References<br />

Apps, C.D., B.N. McLellan, T.A. Kinley, <strong>and</strong> J.P. Flaa. 2001. Scale-dependent habitat selection by<br />

mountain caribou, Columbia Mountains, British Columbia. Journal of Wildlife Mangagement,<br />

65:65-77.<br />

Bergerud, A.T. 1974. Decline of caribou in North America following settlement. Journal of<br />

Wildlife Management, 38:757-770.<br />

Bergerud, A.T. 1978. The status <strong>and</strong> management of caribou in British Columbia. Report to the<br />

Minister of Recreation <strong>and</strong> Conservation. BC Ministry of Recreation <strong>and</strong> Conservation.<br />

Victoria, BC, Canada.<br />

Bergerud, A.T., H.E. Butler <strong>and</strong> D.R. Miller. 1984. Antipredator tactics of calving caribou:<br />

dispersion in mountains. Canadian Journal of Zoology, 62:1566-1575.<br />

Bergerud, A.T. 1985. Anitpredator tactics of caribou: dispersion along shorelines. Canadian<br />

Journal of Zoology, 63:1324-1329.<br />

Bergerud, A.T. <strong>and</strong> J.P. Elliot. 1986. Dynamics of caribou <strong>and</strong> wolves in northern British Columbia.<br />

Canadian Journal of Zoology, 64:1515-1529.<br />

Bradshaw, C.J.A <strong>and</strong> D.M. Hebert. 1996. Woodl<strong>and</strong> caribou population decline in Alberta: fact or<br />

fiction Rangifer, Special Issue No. 9:223-233.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


L.R.P. Williamson et al. 2010. Anthropogenic l<strong>and</strong>scape changes <strong>and</strong> the conservation of woodl<strong>and</strong> caribou<br />

169<br />

Bradshaw, C.J.A, S. Boutin, <strong>and</strong> D.M. Hebert. 1997. Effects of petroleum exploration on woodl<strong>and</strong><br />

caribou in northeastern Alberta. Journal of Wildlife Management, 61:1127-1133.<br />

Culling, D.E., B.A. Culling, T.J. Raabis, <strong>and</strong> A.C. Creagh. 2006. Ecology <strong>and</strong> seasonal habitat<br />

selection of boreal caribou in the Snake-Sahtaneh watershed, British Columbia, 2000-2004.<br />

Report for Canadian <strong>Forest</strong> Products Ltd., Fort Nelson, British Columbia, Canada.<br />

Cumming, H.G. <strong>and</strong> D.B. Beange. 1987. Dispersion <strong>and</strong> movements of woodl<strong>and</strong> caribou near Lake<br />

Nipigon, Ontario. Journal of Wildlife Management, 51:69-79.<br />

Cumming, H. G., D. B. Beange, <strong>and</strong> G. Lavoie. 1996. Habitat partitioning between woodl<strong>and</strong><br />

caribou <strong>and</strong> moose in Ontario: the potential role of shared predation risk. Rangifer, Special<br />

Issue 9:81–94.<br />

Dyer, S.J., J.P. O’Neill, S.M. Wasel <strong>and</strong> S. Boutin. 2001. Avoidance of industrial development by<br />

woodl<strong>and</strong> caribou. Journal of Wildlife Management, 65:531-542.<br />

Dzus, E. 2001. Status of the woodl<strong>and</strong> caribou (Rangifer tar<strong>and</strong>us caribou) in Alberta. Alberta<br />

Environment, Fisheries <strong>and</strong> Wildlife Management Division, <strong>and</strong> Alberta Conservation<br />

Association, Wildlife Status Report No. 30, Edmonton, Canada.<br />

Fuller, T.K, <strong>and</strong> L.B. Keith. 1981. Woodl<strong>and</strong> caribou population dynamics in northeastern Alberta.<br />

Journal of Wildlife Management, 45:197-213.<br />

Gustine, D.D., K.L. Parker, R.J. Lay, M.P.Gillingham <strong>and</strong> D.C. Heard. 2006a. Calf survival of<br />

woodl<strong>and</strong> caribou in a multi-predator ecosystem. Wildlife Monographs, 165:1-32.<br />

Heard, D.C. <strong>and</strong> K.L. Vagt. 1998. Caribou in British Columbia: a 1996 status report. Rangifer,<br />

Special Issue No. 10:159-172.<br />

Holt, R.D. 1977. Predation, apparent competition <strong>and</strong> the structure of prey communities.<br />

Theoretical Population Biology, 12:197-229.<br />

Hummel, M. <strong>and</strong> J.C. Ray. 2008. Caribou <strong>and</strong> the North. Dundurn Press, Toronto, Canada.<br />

James, A.R.C., <strong>and</strong> A.K. Stuart-Smith. 2000. Distribution of caribou <strong>and</strong> wolves in relation to linear<br />

corridors. Journal of Wildlife Management, 64: 154-159.<br />

Johnson, C.J., K.L. Parker, D.C. Heard, <strong>and</strong> M.P. Gillingham. 2002. Movement parameters of<br />

ungulates <strong>and</strong> scale-specific responses to the environment. Journal of Animal Ecology,<br />

71:225-235.<br />

Johnson, C.J., K.L. Parker, D.C. Heard, <strong>and</strong> D.R. Seip. 2004. Movements, foraging habits, <strong>and</strong><br />

habitat use strategies of northern woodl<strong>and</strong> caribou during winter: Implications for forest<br />

practices in British Columbia. BC Journal of Ecosystems <strong>and</strong> Management, 5:22-35.<br />

Jones, E.S., M.P. Gillingham, D.R. Seip, <strong>and</strong> D.C. Heard. 2007. Comparison of seasonal habitat<br />

selection between threatened woodl<strong>and</strong> caribou ecotypes in central British Columbia.<br />

Rangifer, Special Issue No. 17:111-128.<br />

Kinley, T.A., <strong>and</strong> C.D. Apps. 2001. Mortality patterns in a subpopulation of endangered mountain<br />

caribou. Wildlife Society Bulletin, 29:158-164.<br />

Latham, A.D.M. 2009. Wolf ecology <strong>and</strong> caribou-primary prey-wolf spatial relationships in low<br />

productivity peatl<strong>and</strong> complexes in northeastern Alberta. PhD Dissertation. University of<br />

Alberta, Edmonton, Alberta, Canada.<br />

Leopold, A. S., <strong>and</strong> F.F. Darling. 1953. Wildlife in Alaska. Ronald Press, New York, NY.<br />

McNay, R.S. <strong>and</strong> L. Giguere. 2008. Reassessing the supply of northern caribou seasonal range types<br />

in north-central British Columbia. Wildlife Infometrics Report No. 279. Wildlife Infometrics<br />

Inc., Mackenzie, British Columbia, Canada.<br />

Mountain Caribou Technical Advisory Committee. 2002. A strategy for the recovery of mountain<br />

caribou in British Columbia. British Columbia Ministry of water, L<strong>and</strong> <strong>and</strong> Air Protection,<br />

Victoria, British Columbia, USA.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


L.R.P. Williamson et al. 2010. Anthropogenic l<strong>and</strong>scape changes <strong>and</strong> the conservation of woodl<strong>and</strong> caribou<br />

170<br />

Nitschke, C.R. 2008. The cumulative effects of resource development on biodiversity <strong>and</strong><br />

ecological integrity in the Peace-Moberly region of Northeast British Columbia, Canada.<br />

Biodiversity Conservation, 17:1715-1740.<br />

Rempel, R.S., P.C. Elkie, A.R. Rodgers, <strong>and</strong> M.J. Gluck. 1997. Timber-management <strong>and</strong> naturaldisturbance<br />

effects on moose habitat: l<strong>and</strong>scape evaluation. Journal of Wildlife Management,<br />

61:517-524.<br />

Schaefer, J.A. 2003. Long-term range recession <strong>and</strong> the persistence of caribou in the taiga.<br />

Conservation Biology, 17:1435-1439.<br />

Schaefer, J.A. 2008. Boreal forest caribou. In: Hummel, M. <strong>and</strong> J.C. Ray. Caribou <strong>and</strong> the North.<br />

Dundurn Press, Toronto, Canada: 223-237.<br />

Seip, D.R. 1992. Factors limiting woodl<strong>and</strong> caribou populations <strong>and</strong> their interrelationships with<br />

wolves <strong>and</strong> moose in southeastern British Columbia. Canadian Journal of Zoology, 70:1494-<br />

1503.<br />

Seip, D.R. <strong>and</strong> D.B. Cichowski. 1996. Population ecology of caribou in British Columbia. Rangifer,<br />

Special Issue No. 9:73-80.<br />

Seip, D.R., C.J. Johnson, <strong>and</strong> G.S. Watts. 2007. Displacement of Mountain Caribou from Winter<br />

Habitat by Snowmobiles. Journal of Wildlife Management, 71:1539-1544.<br />

Seip, D.R. <strong>and</strong> B.N. McLellan. 2008. Mountain caribou. In: Hummel, M. <strong>and</strong> J.C. Ray. Caribou<br />

<strong>and</strong> the North. Dundurn Press, Toronto, Canada: 240-255.<br />

Smith, K.G., E.J. Smith, D. Hobson, T.C. Sorensen, <strong>and</strong> D. Hervieux. 2000. Winter distribution of<br />

woodl<strong>and</strong> caribou in relation to clear-cut logging in west-central Alberta. Canadian Journal<br />

of Zoology, 78:1433-1440.<br />

Sorensen, T., P.D. McLoughlin, D. Hervieux, E. Dzus, J. Nolan, B. Wynes <strong>and</strong> S. Boutin. 2008.<br />

Determining sustainable levels of cumulative effects for boreal caribou. Journal of Wildlife<br />

Management, 72:900-905.<br />

Spalding, D.J. 1990. The early history of moose (Alces alces): distribution <strong>and</strong> relative abundance<br />

in British Columbia. Contributions to Natural Science Number 11. Royal British Columbia<br />

Museum, Victoria, British Columbia, Canada.<br />

Stephenson RO. 1982. Nunamiut Eskimos, wildlife biologists, <strong>and</strong> wolves. Pages 434–439 in<br />

Harrington FH, Pacquet P.C. (Eds.). Wolves of the World. Park Ridge (NJ): Noyes.<br />

Stevenson, S.K., <strong>and</strong> D.F. Hatler. 1985. Woodl<strong>and</strong> caribou <strong>and</strong> their habitat in southern <strong>and</strong> central<br />

British Columbia. Volume 1. L<strong>and</strong> Management Report No. 23. BC Ministry of <strong>Forest</strong>s.<br />

Victoria, British Columbia, Canada.<br />

Stuart-Smith, K.A., C.J.A. Bradshaw, S. Boutin, D.M. Hebert, <strong>and</strong> A.B.Rippin. 1997. Woodl<strong>and</strong><br />

caribou relative to l<strong>and</strong>scape patterns in northeastern Alberta. Journal of Wildlife<br />

Management, 61:622-633.<br />

Terry, E.L., <strong>and</strong> Wood, M.D. 1999. Seasonal movements <strong>and</strong> habitat selection by woodl<strong>and</strong> caribou<br />

in the Wolverine herd, northcentral B.C. Phase 2: 1994–1997. Rep. No. 204, Peace/Williston<br />

Fish <strong>and</strong> Wildlife Compensation Program, Prince George, British Columbia, Canada.<br />

Vors, L.S., J.A. Schaeffer, B.A. Pond, A.R. Rodgers, <strong>and</strong> B.R. Patterson. 2007. Woodl<strong>and</strong> caribou<br />

extirpation <strong>and</strong> anthropogenic l<strong>and</strong>scape disturbance in Ontario. Journal of Wildlife<br />

Management, 71:1249-1256.<br />

Vors, L.S. <strong>and</strong> M.S. Boyce. 2009. <strong>Global</strong> declines of caribou <strong>and</strong> reindeer. <strong>Global</strong> <strong>Change</strong> Biology,<br />

15:2626-2633.<br />

Wittmer, H.U. 2004. Mechanisms underlying the decline of mountain caribou (Rangifer tar<strong>and</strong>us<br />

caribou) in British Columbia. Ph.D Dissertation. University of British Columbia, Vancouver,<br />

British Columbia, Canada.<br />

Wittmer, H.U., A.R.E. Sinclair, <strong>and</strong> B.N. McLellan. 2005. The role of predation in the decline <strong>and</strong><br />

extirpation of woodl<strong>and</strong> caribou. Oecologia, 144:257-267.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


N. Zurbriggen et al. 2010. Modeling feedbacks between avalanches <strong>and</strong> forests under a changing environment<br />

171<br />

Modeling feedbacks between avalanches <strong>and</strong> forests under a changing<br />

environment in the Swiss Alps<br />

Natalie Zurbriggen 1,3* , Heike Lischke 1 , Peter Bebi 2 & Harald Bugmann 3<br />

1 L<strong>and</strong> Use Dynamics, Swiss Federal Research Institute WSL, Birmensdorf, Switzerl<strong>and</strong><br />

2 Ecosystem Boundaries, WSL Institute for Snow <strong>and</strong> Avalanche Research SLF, Davos,<br />

Switzerl<strong>and</strong><br />

3 <strong>Forest</strong> Ecology, Institute of Terrestrial Ecosystems ITES, ETH Zurich, Switzerl<strong>and</strong><br />

Abstract<br />

<strong>Forest</strong>-avalanche feedbacks are important forces shaping subalpine treeline ecotones. Under<br />

rapid <strong>and</strong> drastic environmental change, these feedbacks can develop unexpected dynamics.<br />

Merging forest <strong>and</strong> avalanche models provides a tool to analyze such feedbacks, <strong>and</strong> to envision<br />

management scenarios for protection forests. We modeled <strong>and</strong> analyzed these feedbacks for<br />

Davos (Switzerl<strong>and</strong>), where l<strong>and</strong> ab<strong>and</strong>onment <strong>and</strong> temperature change are the main drivers of<br />

environmental change. We developed a spatially explicit model for avalanche release zones<br />

inside forests, based on slope steepness, crown coverage, gap size, <strong>and</strong> forest type, <strong>and</strong><br />

incorporated it into the forest-l<strong>and</strong>scape model TreeMig. Preliminary results reveal that (1)<br />

TreeMig is sensitive to the proposed changes in disturbance simulation, (2) a simplified<br />

vegetation model shows feedback effects with an avalanche subroutine, <strong>and</strong> (3) the intermediate<br />

disturbance hypothesis partly applies to avalanche influence on forests. The merged model<br />

TreeMig-Av is presented <strong>and</strong> scenarios of forest-avalanche interaction under environmental<br />

change are interpreted.<br />

Keywords: treeline dynamics, avalanches, environmental change, feedback effects, forest<br />

l<strong>and</strong>scape models<br />

1. Introduction<br />

In subalpine treeline ecotones, avalanches have a strong influence on vegetation dynamics. The<br />

influence of the vegetation on avalanche dynamics, however, is often modeled as a binary<br />

decision variable excluding potential avalanche release in forested areas. Yet it is well known<br />

that avalanches can initiate inside forests, especially in sparsely vegetated forests near the<br />

treeline. So far, there have been only few studies analyzing the interactions between forests on<br />

avalanches, with focus on long-term forest dynamics (Cordonnier 2008). Modeling avalanches<br />

in a forest-l<strong>and</strong>scape model should provide a tool to analyze the future development of forests,<br />

avalanches, <strong>and</strong> their feedbacks, under changing environmental influences. Our goal is to<br />

develop a simplified spatially explicit avalanche module, <strong>and</strong> to implement it in TreeMig<br />

(Lischke et al. 2006), a forest-l<strong>and</strong>scape model based on the gap model ForClim (Bugmann<br />

1994). One of the main goals is to analyze simulations of forests interacting with l<strong>and</strong> use<br />

change <strong>and</strong> climate change, under the influence of avalanches as spatially connected<br />

disturbances. Previous versions of TreeMig accounted only for single-cell disturbances. Output<br />

variables of interest include changes in biomass, species composition, structural composition,<br />

avalanche occurrence, etc. Simulating the protection function of forests against avalanches is a<br />

* Corresponding author. Tel.: 0041 44 739 2817 - Fax: 0041 44 739 2215<br />

Email address: natalie.zurbriggen@wsl.ch<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Zurbriggen et al. 2010. Modeling feedbacks between avalanches <strong>and</strong> forests under a changing environment<br />

172<br />

new option in TreeMig, <strong>and</strong> is the main ecosystem service simulated here, using l<strong>and</strong> use<br />

change scenarios in combination with climate change scenarios as input.<br />

Snow avalanches are a key disturbance in forest ecosystems in the Swiss Alps, <strong>and</strong> their<br />

interactions with forest dynamics, such as mortality <strong>and</strong> regrowth, have strong influences on<br />

subalpine forests <strong>and</strong> their ecosystem services. These feedback effects are especially important<br />

for the location, spatial patterns, <strong>and</strong> structure of upper alpine timberlines. Avalanches can start<br />

both above <strong>and</strong> below timberlines, <strong>and</strong> feedback effects can differ between the two initiation<br />

types. While avalanches starting far above timberlines can destroy forests below without being<br />

influenced much, avalanches starting below timberlines can be strongly modified by the<br />

vegetation. For the avalanche release probability inside forests, st<strong>and</strong> structure <strong>and</strong> gap size are<br />

important factors in addition to the variables commonly used to predict avalanche release<br />

outside of forests, namely topography, weather, <strong>and</strong> snow conditions (Bebi et al. 2009).<br />

Including the vegetation-based variables in an avalanche release module leads to feedback<br />

effects between avalanche release probability <strong>and</strong> the mortality <strong>and</strong> regenerative success of<br />

forests. Yet how these interactions will be affected under scenarios of changing snowpack <strong>and</strong><br />

forest cover is less clear. Merging forest <strong>and</strong> avalanche models provides a tool to analyze such<br />

feedbacks, <strong>and</strong> to envision management scenarios for protection forests under climate <strong>and</strong> l<strong>and</strong>use<br />

change.<br />

Due to the spatial connectivity of avalanches, we expect this disturbance type to have a stronger<br />

influence on the spatial pattern at treelines than the single-cell disturbances previously simulated<br />

in TreeMig. Avalanches are expected to depress treelines to lower altitudes than climate would<br />

allow, <strong>and</strong> to increase fragmentation of the vegetation (Patten et al. 1994), but there are only<br />

few attempts at including avalanches in forest-l<strong>and</strong>scape models (Cordonnier et al. 2008). In<br />

TreeMig, modeled influences on treelines do not differ from influences on lowl<strong>and</strong> areas, except<br />

indirectly via climatic variables <strong>and</strong> differences in seed availability. Most allometries or<br />

processes, such as growth curves, competition for light <strong>and</strong> space, dispersal, or disturbances, are<br />

modeled the same way for all areas. We plan to answer the question of how treelines will<br />

develop under environmental change, <strong>and</strong> how their spatial patterns are influenced by<br />

avalanches, by using TreeMig as a tool for scenario development. To achieve this, the model<br />

needs to be adapted specifically to include more treeline-relevant factors, <strong>and</strong> to allow for<br />

different processes in different simulation areas. Both growth- <strong>and</strong> size-related allometries for<br />

treelines are improved, <strong>and</strong> a process-based feedback loop between avalanches <strong>and</strong> forest<br />

regeneration is implemented.<br />

2. Methods<br />

2.1 Avalanche modeling<br />

The avalanche module is divided into avalanche release <strong>and</strong> avalanche flow, each further<br />

divided into the respective process inside or outside of forested areas. Three of the resulting four<br />

components (release outside forest, flow inside forest, <strong>and</strong> flow outside forest) are based on the<br />

avalanche model RAMMS (Rapid Mass Movements; Christen et al. 2008). The fourth<br />

component, avalanche release inside forested areas, was built as a probabilistic module based on<br />

regression analysis of historical avalanche data (1985-90) <strong>and</strong> related forest data (Meyer-Grass<br />

<strong>and</strong> Schneebeli 1992), according to the method used by Bebi et al. (2001). Using only variables<br />

that can be calculated from TreeMig output, we performed a Generalized Linear Model (GLM)<br />

on the historical avalanche data.<br />

The GLM was used to estimate the dependence of the binary avalanche release variable on<br />

predictor variables describing the vegetation, in addition to the geophysical <strong>and</strong> meteorological<br />

variables often used for avalanche release prediction outside forested areas (e.g. topography,<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Zurbriggen et al. 2010. Modeling feedbacks between avalanches <strong>and</strong> forests under a changing environment<br />

173<br />

snow profile, temperature). The distinct weather conditions within a few days of the event,<br />

which in reality strongly influence avalanche release probability, were not included due to the<br />

larger time scale of TreeMig (1 year steps). To be compatible with TreeMig, the avalanche<br />

module also uses one year steps. A r<strong>and</strong>om component was added to the probabilistic release<br />

module instead of explicit weather related variables, to simulate short-term (within-year)<br />

weather variability. The historical avalanche <strong>and</strong> forest data used as input was stratified into<br />

coniferous <strong>and</strong> broadleaf forest, <strong>and</strong> the GLM is run separately for each forest type.<br />

The selection criteria to include variables in the GLM were (a) significant improvement of<br />

variance explained, (b) sensitivity of at least some of the variables to climatic or l<strong>and</strong> use<br />

change, <strong>and</strong> (c) the possibility to calculate or estimate the variables from TreeMig output. To<br />

implement the equation resulting from the GLM in the avalanche module, the values for slope<br />

angle were taken from the Swiss digital elevation model (DEM) at 25m resolution (Swiss<br />

Federal Office of Topography), <strong>and</strong> maximum gap size <strong>and</strong> other forest-related variables were<br />

calculated allometrically from TreeMig output. For example, crown projection could be<br />

calculated from TreeMig output using the percent cover equations established by Lischke <strong>and</strong><br />

Zierl (2002).<br />

2.2 Mortality <strong>and</strong> regeneration modeling<br />

The increased mortality of trees st<strong>and</strong>ing in an avalanche path will be modeled based on a metamodel<br />

of RAMMS, which is currently in development. In TreeMig, different model versions<br />

will be compared, using either full mortality of all trees in the avalanche path, or partial<br />

mortality based on tree size <strong>and</strong> species. For preliminary versions of the avalanche subroutine in<br />

the forest-l<strong>and</strong>scape model, mortality will be set to one, <strong>and</strong> adapted later to include the<br />

RAMMS meta-model output.<br />

To simulate regeneration after avalanche disturbances, the dispersal, germination, establishment,<br />

<strong>and</strong> growth subroutines of TreeMig will be used (Lischke et al. 2006). Here, growth is modeled<br />

in height class increments, with a maximum possible growth rate per species, modified by<br />

environmental factors such as light, temperature, precipitation, or local disturbances. As trees<br />

are in competition for light, an increased light availability due to avalanches changes the growth<br />

potential of surviving seedlings or freshly germinated seeds.<br />

TreeMig is set up so that primary succession in the destroyed cells is given by potentially<br />

remaining seeds or seedlings, dispersal distance from mature forest, <strong>and</strong> germination <strong>and</strong><br />

growth response to environmental conditions. Seedlings that survive the avalanches will have<br />

higher growth rates due to the lower competition <strong>and</strong> higher light levels, making seedling shade<br />

tolerance <strong>and</strong> the shading function highly critical. To increase the accuracy of the seedling<br />

growth, we will partition the lowest height class (previously including all individuals 0-1.37m<br />

height) into 4 smaller classes, <strong>and</strong> improve the shading subroutine. Furthermore, the previously<br />

strict allometry between size <strong>and</strong> growth, both simulated in units of height classes, will be<br />

improved by allowing more variability between size <strong>and</strong> growth. This is especially important at<br />

treelines, where individuals are often found to be relatively old, but short, with a relatively high<br />

stem diameter.<br />

2.3 Merging of model components<br />

The four components of avalanche release <strong>and</strong> flow inside <strong>and</strong> outside of forested areas will be<br />

implemented in the forest l<strong>and</strong>scape model TreeMig, <strong>and</strong> analyzed for sensitivity, scaling<br />

effects, <strong>and</strong> model uncertainties. To implement the avalanche module in TreeMig, both the<br />

avalanche submodel <strong>and</strong> the growth <strong>and</strong> regeneration subroutines require careful calibration.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Zurbriggen et al. 2010. Modeling feedbacks between avalanches <strong>and</strong> forests under a changing environment<br />

174<br />

Furthermore, to be able to model avalanches in sufficient detail, the cell size will be reduced<br />

from 100m side length to 25m. To ensure applicability of the forest-l<strong>and</strong>scape model with<br />

reduced cell size, we will compare model runs with both cell side lengths.<br />

The feedback effects will emerge once the avalanche release probability, flow dynamics,<br />

destruction, <strong>and</strong> regeneration subroutines are set up. These will have to be fine-tuned for current<br />

climatic <strong>and</strong> l<strong>and</strong>-use conditions to make sure that there are no runaway effects, which could<br />

develop due to the positive feedback cycle between forests <strong>and</strong> avalanches. Keeping the<br />

environmental conditions constant, we will make sure there is a balance between avalanches <strong>and</strong><br />

forest dynamics, by tuning the avalanche <strong>and</strong> forest parameters involved. Once the feedbacks<br />

are established, we will compare the output to current treeline positions <strong>and</strong> patterns, to estimate<br />

the precision of the model.<br />

In a sensitivity analysis, environmental conditions are again kept constant while disturbance<br />

intensity (frequency <strong>and</strong> magnitude) are varied, <strong>and</strong> the influence on output variables such as<br />

total biomass per area, location <strong>and</strong> pattern of treelines, <strong>and</strong> species composition will be<br />

analyzed. The merged model TreeMig-Av will then be applied to the study area of the Davos<br />

municipality, in the eastern Swiss Alps. The model will be validated against historical avalanche<br />

frequency data <strong>and</strong> compared to the output of other models such as those used for avalanche risk<br />

mapping. IPCC AR4 climatic change scenarios <strong>and</strong> different l<strong>and</strong> use scenarios will then be<br />

used to simulate avalanche release <strong>and</strong> forest cover scenarios for the next centuries, <strong>and</strong> to<br />

analyze how the feedback effects develop under the changing environmental conditions.<br />

3. Preliminary <strong>and</strong> expected results<br />

In the GLM, the final variables used for the estimation of avalanche release inside forests are<br />

similar between the two forest types: While the coniferous forest model equation uses slope<br />

angle, crown projection, <strong>and</strong> maximum gap size, the broadleaf forest equation uses slope angle,<br />

crown projection, <strong>and</strong> proportion of coniferous trees instead of the gap size.<br />

A sensitivity analysis in the current version of TreeMig, with single cell disturbances, showed<br />

that it is sensitive to changes in disturbance frequency <strong>and</strong> magnitude. <strong>Change</strong>s in these<br />

variables led to changes not only in total biomass, but also species composition <strong>and</strong> structural<br />

diversity in the affected cells. Furthermore, the single cell disturbance regime was compared to<br />

spatially connected disturbances on a transect, which confirmed the expected difference in<br />

regeneration patterns after the two disturbance types. The main feature of the spatially<br />

connected disturbance type is the potentially larger distance of a recently disturbed cell to the<br />

nearest vegetated cell, from where seeds may disperse into the disturbed cell. Regeneration<br />

dynamics therefore strongly depend on accurate simulation of dispersal kernels <strong>and</strong> size of the<br />

disturbed area.<br />

The Intermediate Disturbance Hypothesis (IDH) was shown to apply to TreeMig's disturbance<br />

rates at the single cell level, <strong>and</strong> for connected disturbances along the one-dimensional transect,<br />

<strong>and</strong> is expected to also apply to spatially connected disturbances in two-dimensional space.<br />

Highest species diversity was found at intermediate disturbance frequency <strong>and</strong> magnitude. The<br />

exact location of the peak diversity along the two axes, however, was influenced by altitude, i.e.<br />

by limitations given by climatic variables. For example, where growth <strong>and</strong> survival is limited by<br />

climate, a relatively high disturbance frequency could lead to the disappearance of the forest<br />

altogether. Here, the peak of diversity would shift to lower values of disturbance frequency. The<br />

IDH applies to species composition but not necessarily to forest structure (size distribution), as<br />

avalanches can increase the number of smaller trees relative to the larger trees. Applicability of<br />

IDH to structural diversity will further be studied using the adapted version of TreeMig.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Zurbriggen et al. 2010. Modeling feedbacks between avalanches <strong>and</strong> forests under a changing environment<br />

175<br />

Potential release areas (PRAs) were simulated in the avalanche release subroutine as results of<br />

the GLM equations, <strong>and</strong> are influenced by the avalanche flow dynamics <strong>and</strong> the subsequent<br />

forest regeneration. The forest has a strong influence in reducing PRA occurrence, but we found<br />

that the forest structure is just as important. Sparsely vegetated areas, which under some<br />

definitions may still be classified as forests, can only reduce PRAs up to a certain point, but can<br />

not avoid avalanches altogether. Furthermore, a forest cannot stop an avalanche once it reaches<br />

full speed <strong>and</strong> strength, which can happen for example when avalanches are initiated far above<br />

forests.<br />

Avalanches are expected to have a stronger influence on vegetation at higher altitudes, due to<br />

the increased release probability at higher altitudes, <strong>and</strong> the downhill flow direction. <strong>Forest</strong>s at<br />

higher altitudes should therefore experience more disturbances by avalanches. As these forests<br />

are also climatically more limited in regeneration, they should be more sensitive to disturbances.<br />

Regarding the IDH, diversities of forests at high altitudes are strongly influenced by<br />

disturbances, <strong>and</strong> may not be ably to maintain the same diversity without avalanches (Rixen et<br />

al. 2007). However, climate change could further shift the maxima along the axis of disturbance<br />

intensities, <strong>and</strong> therefore also influence forest diversity <strong>and</strong> structure.<br />

4. Conclusions<br />

Including an avalanche module in a forest-l<strong>and</strong>scape model provides a useful tool, <strong>and</strong> improves<br />

model predictions for the avalanche protection function. Compared to models that include<br />

vegetation only as a binary variable, our merged model is able to predict PRAs relatively well,<br />

even though it does not consider the same physical or short-term weather variables most<br />

avalanches models use. The purpose of our merged model is not to explicitly account for single<br />

avalanche events, but to generate scenarios of trends for forest-avalanche feedbacks, forest<br />

protection abilities, <strong>and</strong> their development over time under changing environment. We expect<br />

that the proposed changes in TreeMig will lead to an improved representation of subalpine<br />

forest <strong>and</strong> treeline ecotones for different climate change scenarios, <strong>and</strong> therefore to a better<br />

judgment of how effective different forests are in terms of avalanche protection under various<br />

environmental scenarios. This should then contribute to improved long-term decision support<br />

for forest management of the study area.<br />

References<br />

Bebi, P., F. Kienast, <strong>and</strong> W. Schonenberger. 2001. Assessing structures in mountain forests as a<br />

basis for investigating the forests' dynamics <strong>and</strong> protective function. <strong>Forest</strong> Ecology <strong>and</strong><br />

Management 145: 3-14.<br />

Bebi, P., D. Kulakowski, <strong>and</strong> C. Rixen. 2009. Snow avalanche disturbances in forest ecosystems<br />

- State of research <strong>and</strong> implications for management. <strong>Forest</strong> Ecology <strong>and</strong> Management<br />

257: 1883-1892.<br />

Bugmann, H. 1994. On the ecology of mountainous forests in a changing climate: a simulation<br />

study. PhD Thesis. Dissertation Nr 10638, ETH Zurich.<br />

Christen, M., P. Bartelt, J. Kowalski, <strong>and</strong> L. Stoffel. 2008. Calculation of dense snow<br />

avalanches in three-dimensional terrain with the numerical simulation program<br />

RAMMS. Online documentation.<br />

http://ramms.slf.ch/ramms/images/stories/rammsissw08.pdf (last accessed March 2010)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Zurbriggen et al. 2010. Modeling feedbacks between avalanches <strong>and</strong> forests under a changing environment<br />

176<br />

Cordonnier, T., B. Courbaud, F. Berger, <strong>and</strong> A. Franc. 2008. Permanence of resilience <strong>and</strong><br />

protection efficiency in mountain Norway spruce forest st<strong>and</strong>s: A simulation study. <strong>Forest</strong><br />

Ecology <strong>and</strong> Management 256: 347-354.<br />

Lischke, H. <strong>and</strong> B. Zierl. 2002. Feedback between structured vegetation <strong>and</strong> soil water in a<br />

changing climate: A simulation study. Pages 349-377 Climate <strong>Change</strong>: Implications for<br />

the Hydrological Cycle <strong>and</strong> for Water Management. Kluwer Academic Publishers,<br />

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

Lischke, H., N. E. Zimmermann, J. Bolliger, S. Rickebusch, <strong>and</strong> T. J. Loffler. 2006. TreeMig: A<br />

forest-l<strong>and</strong>scape model for simulating spatio-temporal patterns from st<strong>and</strong> to l<strong>and</strong>scape<br />

scale. Ecological Modelling 199: 409-420.<br />

Meyer-Grass, M. <strong>and</strong> M. Schneebeli. 1992. Die Abhängigkeit der Waldlawinen vom St<strong>and</strong>orts-,<br />

Best<strong>and</strong>es- und Schneeverhältnissen. Schutz des Lebensraumes vor Hochwasser, Muren<br />

und Lawinen, Interpraevent 92. 2: 443-455.<br />

Patten, R. S. <strong>and</strong> D. H. Knight. 1994. Snow avalanches <strong>and</strong> vegetation pattern in Cascade-<br />

Canyon, Gr<strong>and</strong>-Teton National Park, Wyoming, USA. Arctic <strong>and</strong> Alpine Research 26:<br />

35-41.<br />

Rixen, C., S. Haag, D. Kulakowski, <strong>and</strong> P. Bebi. 2007. Natural avalanche disturbance shapes<br />

plant diversity <strong>and</strong> species composition in subalpine forest belt. Journal of Vegetation<br />

Science 18: 735-742.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Section 4<br />

Biodiversity conservation <strong>and</strong> planning in changing<br />

l<strong>and</strong>scapes


Akbarzadeh et al. 2010. Ecotourism <strong>and</strong> controlling forest st<strong>and</strong> damages<br />

178<br />

Ecotourism <strong>and</strong> controlling forest st<strong>and</strong> damages<br />

Case study: Varzegan district North West Iran<br />

M. Akbarzadeh 1* , J. Ajalli 1 & E. Kouhgardi 2<br />

1 Islamic Azad University, Myianeh Branch, East Azarbaijan, Iran.<br />

2 Islamic Azad University, Boushehr Branch, Boushehr, Iran.<br />

Abstract<br />

This paper aim to show using ecotourism how can protection the l<strong>and</strong>s, forest st<strong>and</strong>s <strong>and</strong><br />

biodiversity or not In addition to preserving forest ecosystems <strong>and</strong> biodiversity, natural<br />

protected areas in Arasbaran are homel<strong>and</strong>s for people, largely indigenous, who traditionally<br />

base their resource management on a multiple use strategy. We analyzed l<strong>and</strong> use <strong>and</strong> l<strong>and</strong><br />

cover changes in the Varzegan district in the Arasbaran northern west Iran, where Varzegan<br />

recently incorporated ecotourism to their set of economic activities. We evaluated changes in<br />

l<strong>and</strong> use using vegetation maps from 1997 to 2002 based on Enhanced Thematic Mapper<br />

(ETM+), l<strong>and</strong> sat 7 <strong>and</strong> predicted vegetation cover in 2010 by developing a cellular automata<br />

<strong>and</strong> Markovian chains model. So we selected 2 parts at study area for controlling of damages<br />

<strong>and</strong> monitoring of ecotourism effects on it. We used two scenarios to predict l<strong>and</strong> cover in<br />

2010: (a) forest ecosystems <strong>and</strong> biodiversity implemented at the same rate; (b) forest ecosystem<br />

<strong>and</strong> biodiversity decreases due to the growing dem<strong>and</strong> of ecotourism. Both scenarios predict<br />

slight change in the area. Our results provide guidelines for managing natural resources<br />

managements, suggesting that forest st<strong>and</strong> protection, biodiversity conservation <strong>and</strong> ecotourism<br />

are compatible activities.<br />

Key words: Ecotourism, Biodiversity, Varzegan, <strong>Forest</strong> st<strong>and</strong>s.<br />

1. Introduction<br />

Remote sensing techniques have recently recived lots of attentions in agriculture <strong>and</strong> natural<br />

resources. Natural resources <strong>and</strong> environmental conservation need a lot of attention especially<br />

on under devloped countries. Using remote sensing techniques <strong>and</strong> sateliate data for evaluation<br />

of environmental changes is rapidly growing.<br />

Ecotourism industry is of importance in l<strong>and</strong> conservation <strong>and</strong> income. Natural resources,<br />

especially forest ecosystems in Arasbaran have many ancient parameters for ecotourism using.<br />

2. Methodology<br />

2.1 Study area<br />

The study area is located in North West of IRAN <strong>and</strong> the North Eastern of the East Azerbijan<br />

province, which called Arasbaran, is a mountainous area with elevation between 300 <strong>and</strong> 2700<br />

meters above the see level <strong>and</strong> very near to the Caspian Sea.<br />

* Corresponding author. Tel.: +98914 406 6696 - Fax: +98 423 22 400 85<br />

Email address: m.akbarzadeh@m_iau.ac.ir<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Akbarzadeh et al. 2010. Ecotourism <strong>and</strong> controlling forest st<strong>and</strong> damages<br />

179<br />

o<br />

o<br />

o<br />

o<br />

The area is located between 38 .38′<br />

<strong>and</strong> 38 .52′<br />

latitude <strong>and</strong> between 46 .03′<br />

<strong>and</strong> 46 .15′<br />

longitude. It covers diversity of elevation, slope, population <strong>and</strong> l<strong>and</strong> use <strong>and</strong> includes a variety<br />

of Seashore Rivers, etc. There are above 785 plant species in this forest that 97 species of them<br />

are woody (Sagheb et al, 2004).<br />

Arasbaran includes 11 basins. Varzegan is one of them, which is our study area. Aras river is in<br />

the Northern boundary of the study area. Varzegan basin includes about 58500 ha, which is 9.40<br />

percent of Arasbaran,s total area.<br />

Temprate front Mediterranean <strong>and</strong> Siberian climate causes accumulation of snow in the crest of<br />

the mountains in winter. The study area is under the influence of the Mediterranean climate <strong>and</strong><br />

also Caspian <strong>and</strong> Caucasus climates. Nearest meterology station to the area is kaleibar station<br />

which is 1300 m, above the sea level. Average rain fall in this station in a 20- year period was<br />

461mm.<br />

45% of the raining was in spring, 23% in autumn, <strong>and</strong> 22% in winter <strong>and</strong> % 10 in summer.<br />

Maximum monthly evaporation is %85 in spring. Average yearly temperature is different so<br />

that it may be up to12 o c , it would be 17 o c in low elevations <strong>and</strong> 5 o c in high elevations.<br />

Average temperature in the warmest days of the year is 24 o c in low l<strong>and</strong> <strong>and</strong> 12 o c in high<br />

l<strong>and</strong>s. Average temperature in the coldest days of the year is 10 o c in low l<strong>and</strong>s <strong>and</strong> − 2<br />

o c in<br />

high l<strong>and</strong>s. Minimum temperature may be up to − 29<br />

o c , in the study area . Based on Due<br />

Martteene method the climate situation in Ilighinehchay Basin is: Semihumid with cold<br />

summers (Akbarzadeh <strong>and</strong> Babie. K 2002) (Sagheb et al, 2007) (Sabeti.H, 1976).<br />

The area is bounded among Ararat, Sah<strong>and</strong> <strong>and</strong> Sabalan mountains <strong>and</strong> their activities <strong>and</strong><br />

tectonical frequency formed collection of lime <strong>and</strong> igneous stones among them marn, shale, tuff<br />

<strong>and</strong> conglomerate can be seen.<br />

This collection belongs to tertiary which has been formed due to high motion alpine mountain.<br />

There are a large amount of basalt <strong>and</strong> tracite among them.<br />

There are lime pans in Ghare Dagh Mountains too.<br />

Surface of l<strong>and</strong>s are covered with brown soils. There are many springs in the study area which<br />

are emerged from the crest of rocky mountain.<br />

Raining mostly is snow in study area which its low evaporation has caused some rivers to<br />

emerge from high l<strong>and</strong> <strong>and</strong> enter to Aras river. The main river is ilghineh chay which is the<br />

same name as the basin.<br />

Quality of waters is so good that can be used for human being, wildlife <strong>and</strong> agriculture without<br />

any limitation (Akbarzadeh, 2007) (Mukhdum, 2000).<br />

2. Methodology<br />

The socioeconomic analysis was aimed at examining current l<strong>and</strong>-use practices <strong>and</strong> their spatial<br />

distribution, as well as the different l<strong>and</strong>-use strategies implemented by households within the<br />

NPA (natural protected areas). This analysis was needed to identify the forces behind<br />

LUCC(l<strong>and</strong> use / cover changes) in VARZEGAN, since l<strong>and</strong>scapes are transformed by natural<br />

resources appropriation <strong>and</strong> implementation of productive activities, in addition to ecological<br />

processes. Our intention with this analysis was to reveal the ecological, economic <strong>and</strong> social<br />

conditions under which the VARZEGAN ecosystem is managed. Socioeconomic data was<br />

collected between September 2008 <strong>and</strong> October2009.<br />

Data collection was done via participant observation, informal <strong>and</strong> semi-structured interviews,<br />

<strong>and</strong> a survey with semi-closed questions.<br />

All the above were conducted with the assistance of a bilingual interpreter who was not from<br />

VARZEGAN <strong>and</strong> whose native <strong>and</strong> second languages were Turkish <strong>and</strong> Persian, respectively.<br />

Data collection was designed to gather both quantitative <strong>and</strong> qualitative information. Semi-<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Akbarzadeh et al. 2010. Ecotourism <strong>and</strong> controlling forest st<strong>and</strong> damages<br />

180<br />

structured interviews were carried out in 29 of the 30 households inside the NPA, <strong>and</strong> were<br />

focused on household activity implementation. Special emphasis was placed on gathering data<br />

on plots (distance to house, surface area, crops, activities implemented <strong>and</strong> labor allocation). An<br />

additional 23 semi-structured interviews were carried out to gather data on household home<br />

gardens (variety of species <strong>and</strong> uses). Informal interviews <strong>and</strong> semi-closed questions were used<br />

to evaluate their views on ecotourism activities in the area. Emphasis was placed on informant<br />

perceptions of the possibility of allocating more time to new economic activities (ecotourism)<br />

instead of traditional ones.<br />

Finally, a semi-structured interview was held with a tourist agency executive interested in<br />

working in VARZEGAN. Future LUCC scenarios were projected using socioeconomic analysis<br />

data, <strong>and</strong> the factor of possible l<strong>and</strong> tenure changes in VARZEGAN.<br />

To build these scenarios, we classified study area into 2 groups: (a) forest ecosystems <strong>and</strong><br />

biodiversity implemented at the same rate; (b) forest ecosystem <strong>and</strong> bio diversity decreases due<br />

to the growing dem<strong>and</strong> of ecotourism.<br />

Characteristic vegetation in the VARZEGAN NPA is medium deciduous forest in different<br />

successional stages, although small patches of seasonal grassl<strong>and</strong> are also present. Vegetation<br />

maps were generated using various tools <strong>and</strong> Spot 5 data, to validate the different described l<strong>and</strong><br />

use <strong>and</strong> vegetation succession categories in VARZEGAN. Initially, a 2002 Etm+ panchromatic<br />

image was interpreted <strong>and</strong> digi- talized. The visual interpretation of l<strong>and</strong> use categories <strong>and</strong><br />

vegetation successional stages, as well as the mapping of the different trails in the NPA were<br />

aided by performing ground verifications <strong>and</strong> by conducting participatory mapping exercises<br />

with local inhabitants. They have a vast knowledge of plant species, forest successional stages<br />

<strong>and</strong> l<strong>and</strong>-use history which could be translated into spatial references given the scale at which<br />

we were working at. By these processes were defined the following l<strong>and</strong> use <strong>and</strong> vegetation<br />

successional stages categories: agricultural units ( 0–1 year); four forest successional stages (2–<br />

7; 8–15; 16–29; <strong>and</strong> 30–50 years); old-growth forest(>50years); Later, a 2005 SPOT image (5m<br />

ground resolution) was interpreted using the same categories for l<strong>and</strong> use <strong>and</strong> vegetation<br />

successional stages defined for the 2002 map. Additionally, for thisimageweusedb<strong>and</strong>s2,<br />

3<strong>and</strong>4todigitalize, <strong>and</strong>differentiate among textures, forms <strong>and</strong> color patterns (IDRISI15,<br />

Ilwis3.0). The successional stage category was assigned correctly in 90% of the sites.<br />

3. Result<br />

1. Tourism in the NPA will increase mainly because of agreements between local inhabitants<br />

<strong>and</strong> one or more tourists.<br />

2. Protection of area is seen more strongly, because degradation rate in these areas is lower<br />

than the others.<br />

In the first projection scenario (Scenario 1), l<strong>and</strong> use cover maps for 1999 <strong>and</strong> 2003 were<br />

projected to 2007 <strong>and</strong> then to2010under the assumption that the change observed between1999<br />

<strong>and</strong> 2003 would occur at the same pace from 2003 to2007.<br />

The second scenario (Scenario 2) was run in which the probabilities of observed change were<br />

altered according to socioeconomic analysis results <strong>and</strong> current l<strong>and</strong> tenure policy in the NPA.<br />

In Scenario 2 we decreased by 8% the future probability that all succession categories would<br />

change to a new agriculture.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Akbarzadeh et al. 2010. Ecotourism <strong>and</strong> controlling forest st<strong>and</strong> damages<br />

181<br />

Figure 1: Map of Study Area group A: Dark Green B: Light Green<br />

3. Discussion<br />

This hypothetical reduction was based on the results of the socioeconomic analyses, which<br />

suggested a tendency for households to devote increasingly more time to ecotourism <strong>and</strong> less to<br />

agriculture <strong>and</strong> traditional activities, as well as a tendency of local inhabitants to emigrate <strong>and</strong><br />

ab<strong>and</strong>on their agriculture.<br />

References<br />

Akbarzadeh, M., 2007. Using RS <strong>and</strong> GIS for L<strong>and</strong> evaluation in Arasbaran Iran. Ph.D. thesis,<br />

IAU, Science <strong>and</strong> Research University, Tehran, Iran.<br />

Campbell, L.M., Smith, C., 2006. What makes them pay Values of volunteer tourists working<br />

for sea turtle conservation. Environmental Management, 38: 84–98.<br />

Clifton, J., Benson, A., 2006. Planning for sustainable ecotourism: the case for research<br />

ecotourism in developing country destinations. Journal of Sustainable Tourism, 14: 238–<br />

254.<br />

Collar, N.J., 1998. Information <strong>and</strong> ignorance concerning the word’s parrots: an index for<br />

twenty-first century research <strong>and</strong> conservation. Papageienkunde, 2: 201–235.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Carvalho-Santos 2010. Fine-scale mapping of High Nature Value farml<strong>and</strong>s<br />

182<br />

Fine-scale mapping of High Nature Value farml<strong>and</strong>s: novel<br />

approaches to improve the management of rural biodiversity <strong>and</strong><br />

ecosystem services<br />

Cláudia Carvalho-Santos 1* , Rob Jongman 2 , Joaquim Alonso 3 & João Honrado 1<br />

1 Departamento de Biologia, Faculdade de Ciências & CIBIO-Centro de Investigação<br />

em Biodiversidade e Recursos Genéticos, Universidade do Porto, Edifício FC4, Rua do<br />

Campo Alegre, S/N 4169-007 Porto, Portugal<br />

2 Alterra, Wageningen UR, PO Box 47, 6700 AA Wageningen, The Netherl<strong>and</strong>s<br />

3<br />

<strong>Escola</strong> <strong>Superior</strong> Agrária, Instituto Politécnico de Viana do Castelo, 4990-706 Ponte de<br />

Lima<br />

Abstract<br />

High Nature Value farml<strong>and</strong>s (HNVf) are defined as rural l<strong>and</strong>s characterized by high<br />

levels of biodiversity <strong>and</strong> extensive farming practices. These farml<strong>and</strong>s are also known<br />

to provide important ecosystems services, such as food production, pollination, water<br />

purification <strong>and</strong> l<strong>and</strong>scape recreation. Recently, this concept has been introduced in<br />

Rural Development Programmes related to biodiversity preservation in traditional<br />

agricultural l<strong>and</strong>scapes. However, there are no specific rules concerning the practical<br />

use of the concept, particularly on the identification of potential HNVf areas at a local<br />

scale. This application becomes important for farml<strong>and</strong> biodiversity protection in the<br />

context of multi-scale agricultural development.<br />

We present a novel approach for HNVf mapping, which provides an improved local<br />

discrimination of farml<strong>and</strong>s according to their contribution for the conservation of rural<br />

biodiversity <strong>and</strong> ecosystem services. Our approach is based on a multi-criteria valuation<br />

of habitat types based on the national l<strong>and</strong> cover map <strong>and</strong> agrarian censuses. It is<br />

considered applicable in other EU countries since comparable datasets are usually<br />

available. This methodology is also expected to provide the backbone of a st<strong>and</strong>ard,<br />

cost-effective methodology for HNVf monitoring, with an emphasis on the impacts of<br />

l<strong>and</strong> use change on species, habitats <strong>and</strong> l<strong>and</strong>scape function.<br />

Keywords: ecosystem services, local scale, HNVf, mapping, rural biodiversity<br />

1. Introduction<br />

Biodiversity is an important product of agricultural l<strong>and</strong>scapes, but in many European farml<strong>and</strong>s<br />

species richness has been declining (Billeter et al. 2008). Furthermore, research <strong>and</strong> policy on<br />

biodiversity conservation <strong>and</strong> agricultural management have not progressed very well (Moonen<br />

<strong>and</strong> Bàrberi 2008). Since rural l<strong>and</strong>scapes are dominant in most European countries <strong>and</strong> the<br />

European Union (EU) has established ambitious goals concerning the halting of biodiversity<br />

loss (Pereira <strong>and</strong> Cooper 2006; EEA 2006a, 2006b; Fontaine 2007), it is imperative to establish<br />

* Corresponding author. Tel.: (+351) 220402000<br />

Email address: claudiasantos.malta@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Carvalho-Santos 2010. Fine-scale mapping of High Nature Value farml<strong>and</strong>s<br />

183<br />

sound frameworks to monitor agricultural impacts on biodiversity, by selecting the best general<br />

indicators (EASAC 2005; EEA 2005, 2006b, 2007) <strong>and</strong> at the same time paying attention to the<br />

specificity of different agro-ecosystems. It is important to underst<strong>and</strong> the relationships between<br />

l<strong>and</strong>scape, biodiversity <strong>and</strong> l<strong>and</strong> use to manage the l<strong>and</strong> <strong>and</strong> to develop consistent plans for the<br />

future maintenance or enhancement of the current resources (Jongman et al 2006).<br />

More than 50% of Europe’s most highly valued biotopes occur in low intensity farml<strong>and</strong><br />

(Bignal <strong>and</strong> McCracken 1996). Over the last few decades biodiversity losses in farml<strong>and</strong>s were,<br />

in great extent, due to large scale rationalization <strong>and</strong> intensification of agricultural production<br />

<strong>and</strong>, on the other h<strong>and</strong>, many marginal <strong>and</strong> extensively farmed areas have either been improved<br />

or ab<strong>and</strong>oned, both resulting in the reduction on habitats <strong>and</strong> species diversity (EEA 2004).<br />

2. High Nature Value Farml<strong>and</strong><br />

Among the many initiatives to prevent biodiversity decline, the identification <strong>and</strong> mapping of<br />

High Nature Value Farml<strong>and</strong>s (HNVf, low-intensity traditional agricultural areas, such as the<br />

montados in Portugal) is surely one of the most valuable (Andersen et al. 2003; EEA 2004;<br />

Paracchini et al. 2006; Cooper et al. 2007; Poux <strong>and</strong> Ramain 2009). Besides gathering<br />

information about these areas, a major objective is to take conservation measures to protect<br />

hotspots of biodiversity (EEA 2004).<br />

HNVf is a concept applied on rural l<strong>and</strong>s characterized by the existence of high levels of<br />

biodiversity, <strong>and</strong> by extensive farming practices (EEA 2004). Recently, this concept, originally<br />

introduced by Baldock (Baldock D et al. 1993) as farming systems with low-inputs of chemicals<br />

<strong>and</strong> of management practices, was also adapted to the forestry sector in the framework of Rural<br />

Development Plans (Beaufoy <strong>and</strong> Cooper 2008).<br />

Europe is characterized by unique <strong>and</strong> variable rural l<strong>and</strong>scapes, heritage of many centuries of<br />

cultural <strong>and</strong> natural history (EEA 2004). Many of them can be considered as HNVf. According<br />

to Andersen (Andersen et al. 2003), there are three types of High Nature Value farml<strong>and</strong>:<br />

Type 1: farml<strong>and</strong> with a high proportion of semi-natural vegetation;<br />

Type 2: farml<strong>and</strong> with a mosaic of habitats <strong>and</strong>/or l<strong>and</strong> uses;<br />

Type 3: farml<strong>and</strong> supporting rare species or a high proportion of their European or world<br />

populations.<br />

In Europe (EU 15), about 15-25% of the utilized agricultural area (UAA) is considered as HNV<br />

farml<strong>and</strong>. The majority of this area is located in the Southern Europe, <strong>and</strong> in Portugal the<br />

percentage of HNV farml<strong>and</strong> is estimated at about 37% of the total UAA (EEA 2004).<br />

Another important concept associated with HNVf is HNV farming, used in more recent<br />

documents (Beaufoy <strong>and</strong> Cooper 2008). It refers not only to the l<strong>and</strong> use (farml<strong>and</strong>) but also to<br />

the associated farming management practices. In the context of Rural Development Programs,<br />

the HNV farming indicator is an obligation of the EU states in order to assess whether the<br />

objectives of rural programmes are being achieved under the strategy of Pillar 2 of the CAP<br />

(Beaufoy <strong>and</strong> Cooper 2008). These indicators were developed, not only to describe <strong>and</strong><br />

characterize where HNVf is located, the farml<strong>and</strong> systems <strong>and</strong> practices as well as species <strong>and</strong><br />

habitats of conservation concern (baseline indicators), but also to survey HNVf, contributing to<br />

monitor agricultural impacts on biodiversity (result <strong>and</strong> impact indicators). Member States are<br />

committed to identify <strong>and</strong> maintain HNV farming, <strong>and</strong> it is important for all countries to<br />

identify these systems in order to implement targeted economic support measures (Beaufoy<br />

2009). Ultimately, HNVf associated with high levels of biodiversity can also be related with the<br />

concept of ecosystem services, since in these traditional agricultural areas ecosystems provide a<br />

range of important ecosystem services such as food, water purification, soil formation, <strong>and</strong><br />

recreation.<br />

3. Mapping HNVf across Europe<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Carvalho-Santos 2010. Fine-scale mapping of High Nature Value farml<strong>and</strong>s<br />

184<br />

3.1. Problems with existing methodologies<br />

Ecological, historical <strong>and</strong> cultural differences in farming l<strong>and</strong>scapes among countries require<br />

region-specific rules to identify HNVf. This paper addresses this problematic <strong>and</strong> presets a new<br />

methodology to map HNVf at a local level, considering the importance of this identification to<br />

the improvement of rural natural <strong>and</strong> economic environment.<br />

The st<strong>and</strong>ard procedures for mapping HNVf in Europe include the use of l<strong>and</strong> use data (CLC -<br />

Corine L<strong>and</strong> Cover), with classes based on an Environmental Stratification (Metzger et al.<br />

2005)When available, the methodology also suggests the use of complementary information on<br />

farming practices, altitude <strong>and</strong> latitude, soil quality, climatic condition, steepness of slope at<br />

national level to improve the cartography (Paracchini et al. 2006). However, the resulting maps<br />

cannot be used to draw conclusions on the presence of HNV farml<strong>and</strong> at the local level, but<br />

only at the regional level (Paracchini et al. 2006). In fact, scale is very important when trying to<br />

map HNVf, because different agro-ecological processes operate at different scales that must be<br />

taken into account.<br />

For HNVf identification at the local scale the application of a downscaling exercise using a<br />

bigger scale l<strong>and</strong> use map seemed to be a good option. In Portugal, local-scale l<strong>and</strong> cover/use<br />

analysis is based on the COS products (Portuguese l<strong>and</strong> cover map, 1:25.000), obtained from<br />

the interpretation of aerial photos. However, there is no satisfactory direct relationship between<br />

the CLC <strong>and</strong> COS classifications (table 1), <strong>and</strong> an HNVf map based on the application of the<br />

regional-scale methodology to a COS map would exhibit more than twice the extent of HNVf<br />

area than a map obtained using the CLC dataset. So, differences in l<strong>and</strong> cover classifications in<br />

maps with different scales may result in very different maps for a same area. Overall, this<br />

implies that the methodology for identifying <strong>and</strong> mapping HNVf should be revised in order to<br />

adapt it to multi-scalar exercises.<br />

3.2. Local scale HNVf mapping – proposal of a new methodology<br />

In order to best consider those areas that could be excluded when applying the CLC<br />

methodology, a new refined methodology has been developed to identify HNVf at local scales.<br />

The national l<strong>and</strong> cover dataset (COS) is the central dataset in this novel methodology.<br />

The first step is to define the “total farml<strong>and</strong> area”, considering not only the pure agricultural<br />

<strong>and</strong> agro-forestry areas, but also small patches of neighbouring forest <strong>and</strong> semi-natural areas<br />

spatially <strong>and</strong> functionally linked with cropl<strong>and</strong>s. We defined the maximum areas of 5ha for<br />

forests <strong>and</strong> 1ha for semi-natural areas. Herewith, we are placing the farml<strong>and</strong> not as fragments<br />

with restricted boundaries, but in its context as a continuous place where biodiversity circulates<br />

among habitats.<br />

Even if the methodology considers two different levels of analysis, the patch level <strong>and</strong> the civil<br />

parish level, the final HNVf map should be presented at the less detailed scale (the parish level),<br />

in order not to lose information in the transition among scales. Moreover, as a l<strong>and</strong>scape<br />

concept, HNVf should not be mapped directly at the individual patch of COS, but using context<br />

attributes such as l<strong>and</strong>scape metrics <strong>and</strong> other features of the farming l<strong>and</strong>scape <strong>and</strong> of the<br />

territory itself (Figure 1). L<strong>and</strong>scape attributes include those related to l<strong>and</strong>scape composition<br />

<strong>and</strong> to l<strong>and</strong>scape structure. Available data on farming (from agrarian censuses) were also added<br />

at the parish level, to identify the importance of primary sector of activity in each parish. Finally,<br />

natural value was taken into account, using available data on regional biodiversity <strong>and</strong><br />

ecosystems from a previous project (FCUP 2009).<br />

Mean values were calculated for all parameters <strong>and</strong> for each parish, based on “total farml<strong>and</strong><br />

area”. To isolate any surrogacy among variables a correlation analysis (e.g. using the Spearman<br />

index) should be carried out. Finally, a global HNV score is obtained for each patch or parish by<br />

reclassifying the selected parameters into five classes using equal breaks, <strong>and</strong> then by averaging<br />

their values. The final scale ranges from 1 (low nature value farml<strong>and</strong>) to 5 (high nature value<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Carvalho-Santos 2010. Fine-scale mapping of High Nature Value farml<strong>and</strong>s<br />

185<br />

farml<strong>and</strong>). The objective is thus not to identify HNVf vs. non-HNVf areas, but instead to<br />

provide a hierarchic zoning of nature value. The parish-level HNVf map is obtained from the<br />

area-weighted mean value of farml<strong>and</strong> patches inside the parish, <strong>and</strong> the final score is expressed<br />

without considering the total extent of each parish.<br />

3.3. Testing the new framework in Northern Portugal<br />

The region chosen to test the methodology was the Baixo Tâmega, in the north of Portugal, a<br />

mosaic of different agrarian systems <strong>and</strong> l<strong>and</strong>scapes that have, in most cases, been suffering<br />

ab<strong>and</strong>onment in the last few decades. Nonetheless, there are some areas with more specialized<br />

<strong>and</strong> intensive agricultural systems, mostly related to wine production along the Douro <strong>and</strong><br />

Tâmega valleys. There are also non-cultivated areas, mostly in mountain areas, with seminatural<br />

vegetation associated with extensive grazing.<br />

Due to the regular presence of semi-natural vegetation types, most of these farml<strong>and</strong>s are<br />

classifiable as HNVf areas (Andersen et al. 2003). The final maps (figure 2) provide<br />

complementary views of HNVf distribution in the area. In fact, the patch-level map identifies<br />

HNVf areas in most of the study area except for the eastern mountain areas (where forestry<br />

prevails), but the parish-level map identifies small farml<strong>and</strong> areas in these mountains as the ones<br />

with the highest nature value.<br />

4. Discussion<br />

The concept of HNVf, areas associated with low intensity farming, has become very important<br />

regarding agrobiodiversity protection under the Rural Development Programs. It is already used<br />

in many European countries from different perspectives, <strong>and</strong> is starting to be more <strong>and</strong> more<br />

included in the political agricultural context. This could mean economic support to these areas,<br />

through European financial instruments.<br />

L<strong>and</strong> cover, farming characteristics <strong>and</strong> species distribution data are the central datasets in the<br />

common approach to the identification of HNVf at European <strong>and</strong> national levels. The<br />

availability <strong>and</strong> the quality of farming <strong>and</strong> species datasets is a recurrent problem. In fact,<br />

mapping methodologies using l<strong>and</strong> cover datasets at different scales can provide very distinct<br />

HNVf maps.<br />

A new refined methodology based on l<strong>and</strong> cover data, l<strong>and</strong>scape features, farming attributes <strong>and</strong><br />

natural value/conservation data was designed to map HNVf at a local scale. The use of datasets<br />

on nature value including information on the valuation of ecosystem services inferred from<br />

l<strong>and</strong>-use dataset is considered an advantage. In the literature, HNVf is stressed to promote<br />

biodiversity in agroecosystems. In our novel methodology, we suggest a stronger emphasis on a<br />

l<strong>and</strong>scape perspective <strong>and</strong> on the ecosystem services provided by semi-natural areas close to<br />

cropl<strong>and</strong>.<br />

This methodology appears as an important instrument in the identification of HNVf areas to<br />

support policy implementation in the framework of agrobiodiversity protection. It can be used<br />

either spatially, comparing the extent of potential HNVf areas among different regions, or<br />

temporally, comparing changes in extent of HNVf in one region at different times as a<br />

monitoring effort. Additionally, we expect with future research to check the possibility to adapt<br />

this methodology in other EU countries, where local l<strong>and</strong> cover datasets are usually available.<br />

References<br />

Andersen E., Baldock D., Bennet H., Beaufoy G., Bignal E., Brouwer F., Elbersen B., Eiden G.,<br />

Godeschalk F., Jones G., MaCracken D., Nieuwenhuizen W., Eupen M., Hennekens S.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Carvalho-Santos 2010. Fine-scale mapping of High Nature Value farml<strong>and</strong>s<br />

186<br />

& Zervas G. 2003. Developing a Hight Nature Value Farming area indicator. In. EEA<br />

Copenhagen.<br />

Baldock D, Beaufoy G., Bennett G. & Clark J. 1993. Nature Conservation <strong>and</strong> new direction in<br />

the common agricultural policy. In. IEEP London.<br />

Beaufoy G. 2009. HNV Farming - Explaining the concept <strong>and</strong> interpreting EU <strong>and</strong> national<br />

policy commitments. In. European Forum on Nature Conservation <strong>and</strong> Pastoralism<br />

London.<br />

Beaufoy G. & Cooper T. 2008. Guidance Document to the Member States on the Application of<br />

the Hight Nature Value Impact Indicator. In: European Evaluation Network for Rural<br />

Development. Institute for European Environmental Policy Brussels.<br />

Bignal E.M. & McCracken D.I. 1996. Low-intensity farming systems in the conservation of the<br />

countryside. Journal of Applied Ecology, 33, 413-424.<br />

Billeter R., LiiRA J., Bailey D., Bugter B., Arens P., Augenstein I., Aviron S., Baudry J.,<br />

Bukacek R., Burel F., Cerny M., De Blust G., De Cock R., Diekötter T., Dietz H.,<br />

Dirksen J., Dormann C., Durka W., Frenzel M., Hamersky R., Hendrickx F., Herzog F.,<br />

Klotz S., Koolstra B., Lausch A., Le Coeur D., Maelfait J.P., Opdam P., Roubalova M.,<br />

Schermann A., Schmidt T., Schweiger O., Smulders M.J.M., Speelmans M., Simova P.,<br />

Verboom J., van Wingerden W., Zobel M. & Edwards P.J. 2008. Indicators for<br />

biodiversity in agricultural lansdcapes: a pan-European study. Journal of Applied<br />

Ecology, 45, 141-150.<br />

Cooper T., Arblaster K., Baldock D. & Beaufoy G. 2007. Guidance Document to the Member<br />

States on the Application of the HNV Impact Indicator. In. IEEP Brussels.<br />

EASAC 2005. A user's guide to biodiversity indicators. The Royal Society, London.<br />

EEA 2004. Hight Nature Value farml<strong>and</strong> - Characteristics, trends <strong>and</strong> policy challenges. In.<br />

Europen Environment Agency Copenhagen.<br />

EEA 2005. Agriculture <strong>and</strong> environment in EU-15 — the IRENA indicator report. In. European<br />

Environmental Agency Copenhagen.<br />

EEA 2006a. Integration of environment into EU agriculture policy - the IRENA indicator-based<br />

assessment report. In: Copenhaga.<br />

EEA 2006b. Progress towards halting the loss of biodiversity by 2010. In: Copenhagen.<br />

EEA 2007. Halting the loss of biodiversity by 2010: proposal for a first set of indicators to<br />

monitor progress in Europe. In. European Environment Agency Copenhagen.<br />

FCUP I., UTAD 2009. O Património Natural como factor de desenvolvimento e<br />

competitividade territoriais no Baixo Tâmega - o presente e o futuro do património<br />

natural dos concelhos de Amarante, Baião e Marco de Canaveses In. relatório final da<br />

primeira fase Porto.<br />

Fontaine B.e.a. 2007. The European union's 2010 target: Putting rare species in focus.<br />

Biological Conservation, 139, 167-185.<br />

Metzger M., Bunce R., Jongman R., Mucher C. & Watkins J. 2005. A climatic stratification of<br />

the environment of Europe. <strong>Global</strong> Ecology <strong>and</strong> Biogeography, 14, 549-563.<br />

Moonen A.C. & Bàrberi P. 2008. Functional biodiversity: An agroecosystem approach.<br />

Agriculture, Ecosystems <strong>and</strong> Environment, 127, 7-21.<br />

Paracchini M.L., Terres J.T., Petersen J.E. & Hoogenveen Y. 2006. Background Document on<br />

the Methodology for Mapping Hight Nature Value Farml<strong>and</strong> in EU27. In. EEA.<br />

Pereira H. & Cooper H. 2006. Towards the global monitoring of biodiversity change<br />

. Trends in Ecology <strong>and</strong> Evolution, 21, 123-129.<br />

Poux X. & Ramain B. 2009. L' Agriculture à Haute Valeur Naturelle: mieux la (re) connaître<br />

pour mieux l'accompagner. In. European Forum on Nature Conservation <strong>and</strong><br />

Pastoralism<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Carvalho-Santos 2010. Fine-scale mapping of High Nature Value farml<strong>and</strong>s<br />

187<br />

Table 1 – L<strong>and</strong> cover/use classes used to identify farml<strong>and</strong> areas when using wither the CLC or the COS<br />

classifications<br />

CLC – Lusitanian region<br />

Pastures<br />

L<strong>and</strong> principally occupied by agriculture<br />

Agro forestry areas<br />

Moors <strong>and</strong> heathl<strong>and</strong>s<br />

COS<br />

Annual crops associated with permanent crops (orchards, vineyards<br />

<strong>and</strong> olive groves)<br />

Orchards <strong>and</strong> orchards associated with olive groves, vineyards <strong>and</strong><br />

annual crops<br />

Olive groves <strong>and</strong> olive groves associated with orchards, vineyards<br />

<strong>and</strong> annual crops<br />

Vineyard <strong>and</strong> vineyards associated with olive groves, orchards <strong>and</strong><br />

annual crops.<br />

Agro forestry areas<br />

Complex <strong>and</strong> partial cultural systems<br />

Semi-natural areas<br />

Total Farml<strong>and</strong> of Baixo<br />

Tâmega map<br />

HNV<br />

features<br />

L<strong>and</strong>scape composition<br />

(Patch level)<br />

• Naturalness 1<br />

• Naturalness 2<br />

• Patch Richness<br />

Density<br />

L<strong>and</strong>scape structure<br />

(Patch level)<br />

• Patch Size<br />

• Mean Nearest<br />

Neighbor<br />

• Interspersion<br />

Juxtaposition Index<br />

• Mean Proximity<br />

Index<br />

• Mean Shape Index<br />

• Edge Density<br />

• Mean Patch Fractal<br />

Dimension<br />

Farming features<br />

(Parish level)<br />

• Mean size of<br />

holdings<br />

• Mean number of<br />

plots by holding<br />

• Mean size of plots<br />

• Ratio permanent<br />

/annual crops<br />

Natural Value<br />

(Patch level)<br />

• Biological value<br />

• Ecosystem value<br />

• Conservation value<br />

Figure 1 – Groups of parameters included in the new methodology to map HNVf at a local scale<br />

Figure 2 – Patch <strong>and</strong> Parish HNVf map using the new methodology<br />

Figure 2 – Patch <strong>and</strong> Parish HNVf map using the new methodology<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Catalano et al. 2010. Wood macrolichen Lobaria pulmonaria on chestnut tree crops: the case study of Roccamonfina park<br />

188<br />

Wood macrolichen Lobaria pulmonaria on chestnut tree crops: the case<br />

study of Roccamonfina park (Campania region - Italy)<br />

I. Catalano 1* , A. Mingo 1 , A. Migliozzi 1 , S. Sgambato 1 & G.G. Aprile 1<br />

Dip. di Arboricoltura, Botanica e Patologia Vegetale, Università degli Studi di Napoli<br />

Federico II, Italy<br />

Abstract<br />

Integrating at l<strong>and</strong>scape level the information coming from local environment indicators can<br />

help monitoring environmental quality in conservation programs. Lobaria pulmonaria is a<br />

lichen species widely used to evaluate the spatio-temporal continuity of forest cover <strong>and</strong> to<br />

assess environmental quality in areas of high biogeographical interest. In the Lobaria project of<br />

the Società Italiana Lichenologica, wood macrolichens were sampled on Chestnut woods in a<br />

regional park of Campania Region (Italy). A geographical datasets of lichen distribution, l<strong>and</strong><br />

use, topographical <strong>and</strong> climatic characteristic was built in GIS environment. Multivariate<br />

analysis was conducted to highlight the relationships between lichen distribution <strong>and</strong><br />

environmental quality. The results show that the agronomic management practiced in this area<br />

enabled the establishment of stable conditions over time <strong>and</strong> the development of species<br />

indicator of undisturbed areas. A general framework to analyse l<strong>and</strong>scape-level changes over<br />

the area is proposed in this paper.<br />

Keywords: Lobaria, lichens, chestnut, l<strong>and</strong>scape, monitoring.<br />

1. Introduction<br />

No areas was left on our planet that did not experience changes, directly or indirectly connected<br />

to human activities. The assessment <strong>and</strong> monitoring of forest resources for environmental policy<br />

<strong>and</strong> management is recognized to be of prime importance (Loppi et al., 1999). Epiphytic lichens<br />

are an integral component of forest ecosystem <strong>and</strong> represent a characteristic part of the total<br />

biodiversity. Habitat fragmentation <strong>and</strong> other human l<strong>and</strong> use variables, such as urbanization,<br />

intensity of agricultural or pastoral use, <strong>and</strong> forestry management, are increasingly important as<br />

predictors of lichen species distribution (Will-Wolf et al., 2002).<br />

In fact, lichens react to disturbances <strong>and</strong> habitat alterations for several reasons. Firstly, some<br />

lichen species are dependent on favorable microclimatic conditions. Some epiphytic lichens,<br />

particulary the rarer ones, are stenotopic <strong>and</strong> require a long habitat continuity, for example<br />

substrates such as old or large trees (Friedel et al., 2006). Many epiphytic lichens are strongly<br />

affected by forestry practices, particularly logging (Nascimbene et al. 2007). For example,<br />

cyanolichens are considered a guild of species extremely sensitive to intensive forest<br />

management <strong>and</strong> so they are good indicators of forest continuity. This is the case of the flagship<br />

species Lobaria pulmonaria (L.) Hoffm., whose populations was drastically reduced in the<br />

recent decades (Nascimbene et al., 2007).<br />

The aim of this work was to analyze the effects of chestnut tree crops management practices on<br />

Lobaria pulmonaria populations, <strong>and</strong> on lichens diversity in a natural reserve of Southern Italy.<br />

* Corresponding author. Tel.: +390817760104 ext.45<br />

Email address: migliozz@unina.it<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Catalano et al. 2010. Wood macrolichen Lobaria pulmonaria on chestnut tree crops: the case study of Roccamonfina park<br />

189<br />

2. Materials <strong>and</strong> methods<br />

The study was carried out in Roccamonfina’s Park a reserve located in the northen part of<br />

Campania, South Italy. According to the phytoclimatic classification (Nimis & Martellos, 2008),<br />

the area is intermediate between humid sub-Mediterranean <strong>and</strong> humid Mediterranean. The mean<br />

annual temperature is 15.7°C <strong>and</strong> the mean annual rainfall 950 mm. <strong>Forest</strong>s in this area are<br />

dominated by chestnut tree crops (Castanea sativa Miller).<br />

The monitoring study was carried out between March <strong>and</strong> July 2008. Sixteen 30x30 m plots<br />

were selected, spaced 300-500 m inside four stations differing for forest cover <strong>and</strong> management.<br />

In the center of each plot, 6 trees colonized by Lobaria pulmonaria were sampled <strong>and</strong> georeferenced.<br />

On the selected trees, the presence of all epiphytic lichens was recorded from the<br />

trunk base to a height of 180 cm. Species nomenclature <strong>and</strong> phytoclimatic classification follow<br />

Nimis & Martellos (2008).<br />

In addition we calculated Lobaria pulmonaria area with the aid of 40x40 cm sampling grids<br />

divided into 1x1 cm contiguous quadrants:<br />

%=(total area of Lobaria pulmonaria/ lateral area of trunk)*100<br />

Cartographic processing was carried out using ArcGis 9.3, Ilwis 3.4 <strong>and</strong> Idrisi Kilimangiaro<br />

GIS integrate software. Kriging was the interpolation method used to obtain a geostatistic<br />

correlation of the nearest r<strong>and</strong>omly surveyed values inside the plots <strong>and</strong> to produce an estimate<br />

of minimum least squares variance, as a continuos map. The interpolation procedure, followed a<br />

pattern analysis that allowed to define the range of action of the algorithm (barriers), in order to<br />

minimize disturbance of adjacent areas not subject to survey. Four separate continuos grids,<br />

strictly related to the single plot, were subsequently elaborated to present Lobaria pulmonaria<br />

coverage percentage maps.<br />

Floristic data were analyzed with multivariate statistical methods (Podani 2007), namely a<br />

classification analysis (cluster analysis) <strong>and</strong> an ordination analysis (PCA, Principal Component<br />

Analysis), using SYN-TAX 2000 software. Cluster analysis was performed with group average<br />

link method (upgma) with chord distance for binary data as the dissimilarity coefficient, in order<br />

to identify the classification dendrogram. For ordination analysis a centred PCA was applied,<br />

obtaining the biplot of the species-stations dispersion <strong>and</strong> the screeplot showing the percentage<br />

of variability explained by each of the detected axes.<br />

Ecological indices were calculated to detect local trends related to pH of the substratum, solar<br />

irradiation, aridity, eutrophication <strong>and</strong> poleophoby referring to the Italian Lichen System (Nimis<br />

& Martellos, 2008).<br />

3. Results<br />

A total of 74 lichen species were recorded during the survey on the 96 sampled trees, four of<br />

which were new findings for Campania region (Calicium abietinum Pers., Lobarina<br />

scrobiculata (Scop.) Nyl., Physconia subpulverulenta (Szatala) Poelt v. subpulverulenta,<br />

Platismatia glauca (L.) W. L. Culb. & C. F. Culb). Most of the lichens (83.56%) presented a<br />

coccoid green alga as photobiont, whereas few of them had cyanobacterium (15.07%) or<br />

Trentepohlia (1.37%). Foliose lichens prevailed (50%), followed by crustose (39.18%) <strong>and</strong><br />

fruticose (10.82%).<br />

Concerning phytoclimatic classification, temperate species prevailed (41.25%), divided in<br />

temperate (26.25%), mild temperate (10%) <strong>and</strong> cold temperate (5%). Then followed the suboceanic<br />

species (22.5%), <strong>and</strong> holartic (17.5%). The remaining 18.75% was divided in several<br />

phytoclimatic groups, among which the mediterranean.<br />

Two clusters for the stations (hereafter cluster A <strong>and</strong> B) <strong>and</strong> five for the species (hereafter<br />

cluster 1 to 5) were identified by classification analysis (Figure 1). The PCA (Figure 2)<br />

confirmed the groups identified by the cluster analysis.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Catalano et al. 2010. Wood macrolichen Lobaria pulmonaria on chestnut tree crops: the case study of Roccamonfina park<br />

190<br />

Figure 1 Dendrograms showing the main clusters of species (above) <strong>and</strong> stations (below)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Catalano et al. 2010. Wood macrolichen Lobaria pulmonaria on chestnut tree crops: the case study of Roccamonfina park<br />

191<br />

Figure 2: Biplot showing the main axis of PCA ruled on species <strong>and</strong> stations<br />

Stations<br />

Cluster A includes the stations 1, 2 <strong>and</strong> 4, located at relatively lower elevations. Cluster B<br />

includes four plots located at higher elevation <strong>and</strong> is characterized by the presence of almost<br />

exclusive species such as Hypogymnia physodes, Platismatia glauca, Physconia<br />

subpulverulenta v. subpulverulenta e Parmelina pastillifera, <strong>and</strong> some other species that are<br />

represented with a higher frequency relative to the other cluster, such as Leprocaulon<br />

microscopicum <strong>and</strong> Lecanora chlarotera.<br />

Species<br />

Cluster 1 includes lichen species of the Lobarion pumonariae alliance, such as Lobaria<br />

amplissima var. amplissima (the spesies with the highest frequency in the sampled area),<br />

Lobarina scrobiculata, Degelia plumbea, Peltigera collina, Nephroma laevigatum, Leptogium<br />

cyanescens.<br />

Cluster 2 is dominated by species of the Xanthorion paretinae alliance, with some species of<br />

Lecarion subfuscae: Lecanora clarotera, Lecanora carpinea, Lecidella elaeochroma, Physcia<br />

adscendens, Physconia distorta, Xanthoria parietina, Punctelia subrudecta. These species are<br />

typical of heliophilous, xerophilous, nitrophilous communities distributed in anthropized<br />

environments, where they show a good resistance to pollution.<br />

Cluster 3: Species of Graphidion scriptae: Opegrapha atra, Ochrolechia balcanica, Pertusaria<br />

amara, Pertusaria hymenea. These are communities of crustose pioneer lichens, usually found<br />

in forest environments <strong>and</strong> rarely in anthropized areas, that precede in the succession or<br />

sometimes coexist with the species of Lobarion <strong>and</strong> Parmelion.<br />

Cluster 4: species of Parmelion parlatae <strong>and</strong> Xanthorion parietinae alliances: Flavoparmelia<br />

caperata, Parmotrema perlatum, Hypogymnia physodes, Parmelia sulcata, Parmelina tiliacea.<br />

These are species of mesophilous <strong>and</strong> sub-acidophilous communities, less adapted to dry <strong>and</strong><br />

nitrified environments compared to Xanthorion, but relatively more sensitive to pollution,<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Catalano et al. 2010. Wood macrolichen Lobaria pulmonaria on chestnut tree crops: the case study of Roccamonfina park<br />

192<br />

though in some case they are found in anthropized areas if atmospheric humidity allow their<br />

presence.<br />

Cluster 5 includes several species that may be considered closer to Lobarion pulmonariae than<br />

to Parmelion perlatae or to Xanthorion parietinae.<br />

Cartographic analysis<br />

Four maps of Lobaria percent cover were obtained for the sampled stations. Lobaria cover<br />

values ranged between 0.009 <strong>and</strong> 35%, clustered in nine classes in the maps (Figure 3). The<br />

species did not show very high cover values, except in one of the plots of the first station.<br />

Figure 3: Lobaria sp Cover map elaborated from the surveyed data<br />

4. Discussion<br />

Lichen biodiversity resulted relatively high in the sampled areas, both for the numerical<br />

representation of taxa <strong>and</strong> for their floristic quality, though rich of generalist species adapted to<br />

disturbed areas. Lobaria cover, however, presented relatively low cover values on the studied<br />

area, as it resulted from cartographic analysis. Also the other species of the Lobarion<br />

pulmonariae alliance were not very diffuse on the area. Ecological indices showed that many of<br />

the sampled species are typical of natural or semi-natural habitats, with a hygro-mesophylous<br />

behaviour, preferring high availability of diffuse light, but escaping direct sun radiation. Though<br />

the list of sampled species represents only a small portion of the lichen flora in the studied area,<br />

the marked presence of temperate <strong>and</strong> sub-oceanic species is well correlated with the submountain<br />

character of the area.<br />

A limited number of species characterized by low tolerance to air pollutants prevailed in cluster<br />

A, whereas a greater number of species was found in cluster B, most of which typical of<br />

undisturbed areas or small mountain urban sites. The most frequent species belonging to the<br />

alliance of Lobarion pulmonariae are Lobaria amplissima <strong>and</strong> Lobaria pulmonaria; other<br />

species, such as Lobarina scrobiculata e Peltigera collina, resulted to be rare <strong>and</strong> sporadic,<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Catalano et al. 2010. Wood macrolichen Lobaria pulmonaria on chestnut tree crops: the case study of Roccamonfina park<br />

193<br />

mainly in the northern side of the park, inside the ancient volcanic caldera. These species are<br />

good indicators of wood stability over time. They were mainly found in chestnut forests, rather<br />

than in coppiced oak mixed wood areas (Quercus pubescens, Olea europaea, etc.), more<br />

represented in the western side of the park. Despite l<strong>and</strong>scape fragmentation, the results could<br />

suggest that the agroforestry management carried out over many decades in the north-east of the<br />

park allowed wide forest corridors to get established, which could be an important factor of<br />

ecological continuity. This ensured the conservation of biodiversity due to the local presence of<br />

a large number of vegetative propagules, inducing several colonization events of rare lichen<br />

species over surrounding areas. So a fruit chestnut wood, subjected to a good agronomic <strong>and</strong><br />

forestry management, would give better results in terms of biodiversity than an oak mixed<br />

forest, characterized by a more natural floristic composition but strongly disturbed by frequent<br />

coppice.<br />

The method used to estimate the actual coverage of the monitored species associated with the<br />

cartographic approach, allowed to quantify the species areal extent as monitored on individual<br />

trees, <strong>and</strong> to propose an estimation of the visible cover of the species in the studied area. This<br />

kind of study represents a good starting point for further investigations, specifically designed to<br />

monitoring possible deviations from current conditions <strong>and</strong> focused to assess the effectiveness<br />

of environmental policy actions taken at municipal <strong>and</strong> regional level.<br />

References<br />

Friedel, A., Oheimb, G. v., Dengler, J. & Härdtle, W., 2006. Species diversity <strong>and</strong> species<br />

composition of epiphytic bryophytes <strong>and</strong> lichens – a comparison of managed <strong>and</strong><br />

unmanaged beech forests in NE Germany. Feddes Repertorium, 117: 172-185.<br />

Loppi, S., Bonini, I. & De Dominicis, V., 1999. Epiphytic lichens <strong>and</strong> bryophytes of forest<br />

ecosystem in Tuscany (cental Italy). Cryptogamie, Mycologie, 20 (2): 127-135.<br />

Nascimbene, J., Marini, L., Nimis, P.L., 2007. Influence of forest management on epiphytic<br />

lichens in a temperate beech forest of northern Italy. <strong>Forest</strong> Ecology <strong>and</strong> Management,<br />

247: 43–47.<br />

Nimis, P.L. & Martellos, S., 2008. ITALIC - The Information System on Italian Lichens.<br />

Version 4.0. University of Trieste, Dept. of Biology, IN4.0/1<br />

(http://dbiodbs.univ.trieste.it/).<br />

Podani, J., 2007. Analisi ed esplorazione multivariata dei dati in ecologia e biologia. Liguori<br />

editore, Napoli.<br />

Will-Wolf, S., Scheidegger, C. <strong>and</strong> McCune, B., 2002. Methods for monitoring biodiversity <strong>and</strong><br />

ecosystem function. In: Nimis P.L., Scheidegger C. & Wolseley P. (eds.). Monitoring<br />

with Lichens – Monitoring Lichens. Kluwer Academic Publishers, Dordrecht, The<br />

Netherl<strong>and</strong>s, pp. 147-162.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Extent <strong>and</strong> characteristics of mire habitats in Galicia (NW Iberian Peninsula)<br />

194<br />

Extent <strong>and</strong> characteristics of mire habitats in Galicia (NW Iberian<br />

Peninsula): implications for their conservation <strong>and</strong> management<br />

Ramón Alberto Diaz-Varela, Pablo Ramil-Rego, Manuel Antonio Rodríguez-Guitián &<br />

Carmen Cillero-Castro<br />

1 Universidade de Santiago de Compostela, Spain<br />

Abstract<br />

NW Iberian Peninsula hosts a variety of mire habitats linked to wet environments that contrast<br />

with surrounding regional vegetation. These wetl<strong>and</strong>s are confined to locations meeting certain<br />

azonal conditions, due to local topographies that prevent or delay drainage, to the occurrence of<br />

significant rainfall, or to a combination of both. Despite of their international relevance <strong>and</strong> their<br />

designation as habitats of interest for biodiversity conservation by the EU habitats Directive,<br />

they are threatened mainly by changes in l<strong>and</strong> use, particularly by agriculture<br />

intensification/ab<strong>and</strong>onment <strong>and</strong> the development of wind farms.<br />

In the present work we address the mapping, description <strong>and</strong> analysis of these habitats in a<br />

spatial explicit way in the region of Galicia (NW Iberian Peninsula), in order to support the<br />

development of strategies of planning <strong>and</strong> management. Results showed differences in the<br />

extent, distribution <strong>and</strong> characteristics of the different types of mire habitats with important<br />

implications in their conservation.<br />

Keywords: Mire habitats; NW Iberian Peninsula; spatial distribution; habitat environmental<br />

controls; CHAID<br />

1. Introduction<br />

Mire habitats are azonal wetl<strong>and</strong> ecosystems confined to areas with particular environmental<br />

conditions defined by a positive water balance, a low organic mater decomposition rate <strong>and</strong><br />

where the vegetation remains composition has the potential to form peat (Goodwillie 1980;<br />

Raeymaekers 2000; Rydin <strong>and</strong> Jeglum 2008). Their conservation value has been recognized<br />

internationally by different institutions <strong>and</strong> treaties, as the Ramsar Convention. In the EU<br />

context, most of mire types were included as interest of priority habitats in the Annex II of the<br />

European Directive 92/43/CEE for their inclusion in the European network of protected areas<br />

Natura 2000 (European Commission, 2003). They also host a great amount of species of<br />

bacteriae, protozoans, fungi, algae, lichens, bryophytes, vascular plants, invertebrates <strong>and</strong><br />

vertebrates of interest for biodiversity conservation. Most of them have they optimal or even<br />

exclusive habitats in these environments (Rydin <strong>and</strong> Jeglum 2008). In relation with plant<br />

communities, mire habitats frequently include endemic taxa <strong>and</strong> usually show particular floristic<br />

compositions along geographical <strong>and</strong> altitudinal gradients because of their azonal <strong>and</strong><br />

fragmented distribution.<br />

A key issue in the planning <strong>and</strong> conservation of these habitats is the assessment <strong>and</strong> modelling<br />

of their spatial distribution in relation with key environmental factors as this information might<br />

be used for the optimization of conservation efforts (Wainwright <strong>and</strong> Mulligan 2004). There are<br />

a number of factors that determine the occurrence <strong>and</strong> the type of mire habitats, being the most<br />

important climate, topography <strong>and</strong> nutrient supply (Graniero <strong>and</strong> Price 1999). In this work we<br />

aimed at the exploration of the relationship <strong>and</strong> dependence between different environmental<br />

controls <strong>and</strong> the occurrence of different mire habitats in a sector of the NW of Spain. More<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Extent <strong>and</strong> characteristics of mire habitats in Galicia (NW Iberian Peninsula)<br />

195<br />

specifically, we carried out a data mining procedure to determinate the influence of topography,<br />

climate, lithology <strong>and</strong> distance to sea in the occurrence of four types of mire habitats.<br />

2. Methodology<br />

We conducted our analyses in the region of Galicia, located in the NW of Spain (code NUTS<br />

ES11 of the European Union) <strong>and</strong> covering an area of 29 574 Km2 (cf. fig. 1). Altitudes range<br />

from sea level to about 2000 m a.s.l. showing a contrasted relief. Biogeographically most of the<br />

area is included in the Atlantic Region, but a sector of the SE quadrant of the scene corresponds<br />

with Mediterranean region, being a matter of discussion the exact extent of the Mediterranean<br />

domain in the area (European Environment Agency 2008, Rodríguez Guitián <strong>and</strong> Ramil Rego<br />

2008; Rivas-Martinez <strong>and</strong> Rivas-Saenz 2009). Natural vegetation comprises different deciduous<br />

forest, mostly dominated by Quercus robur, but most of the present day l<strong>and</strong> cover shows an<br />

important degree of human intervention, being frequent afforestations with non native species,<br />

scrubs <strong>and</strong> heatl<strong>and</strong>s along with mosaics of traditional <strong>and</strong> modern agrarian l<strong>and</strong>scapes.<br />

Mire habitats maps were done using different data sources, from national habitat inventories<br />

(http://www.mma.es/portal/secciones/biodiversidad/banco_datos/info_disponible/index_atlas_m<br />

anual_habitats.htm) <strong>and</strong> wetl<strong>and</strong> catalogs (Galician Wetl<strong>and</strong> Inventory,<br />

http://medioambiente.xunta.es/espazosNaturais/humidais/index.htm) to monographic works<br />

(Izco Sevillano et al. 2001) <strong>and</strong> regional maps for protected areas planning (Ramil-Rego <strong>and</strong><br />

Crecente Maseda 2005). All this information was processed in a Geographic Information<br />

System software (ArcGisTM V9.2) <strong>and</strong> converted to different raster layers (one for mire type)<br />

with a cell size of 90 m. Mire classification legend was based on the typology of the Annex I of<br />

the European Habitat Directive (consolided version of the 92/43/CEE Directive) <strong>and</strong> included<br />

four types of mire habitats: Blanket bogs, Cladium mariscus fens, Calcareous fens <strong>and</strong> Tufa<br />

formations, <strong>and</strong> was a compromise between the available input information <strong>and</strong> the categories<br />

of the Directive. Other mire habitats of the Annex I of the directive (i.e. Raised bogs,<br />

Depressions on peat substrates of the Rhynchosporion <strong>and</strong> Transition mires <strong>and</strong> quaking bogs)<br />

were not considered in this analysis as are often included in mosaics of other habitat complexes<br />

(frequently wet heathl<strong>and</strong> or moorl<strong>and</strong>) in the existing cartography.<br />

We considered four groups of explanatory variables (climate, topography, geology <strong>and</strong> distance<br />

to sea) as potential environmental controls for mire habitats (cf. table 1). Climate variables were<br />

obtained from regional climate maps available in literature (Martínez Cortizas <strong>and</strong> Pérez Alberti<br />

1999), where the variables are expressed in intervals of different amplitude labelled as a<br />

numeric scale (cf. table 2). Topographic variables were computed from a Digital Elevation<br />

Model using the ArcGisTM V9.2 software package. Geology map were done by reclassifying<br />

the original classes of geological charts from the Spanish Geological Institute in five classes,<br />

namely sedimentary (sd), granitic (gr), ultrabasic (ub), siliceous metamorphic (mt) <strong>and</strong><br />

calcareous (ca) materials. We also computed the minimum distance to sea as an indicator of the<br />

degree of continentality. All the information layers were converted to raster format with the<br />

same spatial reference <strong>and</strong> resolution as the mire habitat maps. We extracted the explanatory<br />

variables data by means of the spatial overlay of the masks corresponding with each mire type.<br />

The final output for the subsequent analyses was a table with mire type as dependent / grouping<br />

variable <strong>and</strong> several explanatory or independent variables.<br />

In order to asses the effect or explanatory power of these variables, we performed a<br />

classification using the tree based segmentation technique CHAID, acronym of Chi-Squared<br />

Automatic Interaction Detection (Biggs et al. 1991; Kass 1980). CHAID is an exploratory<br />

analysis based in the recursive partitioning of a feature space of several independent or potential<br />

predictors, that themselves might interact, in relation to a dependent or response variable. Both<br />

predictors <strong>and</strong> response variables may be continuous, ordinal or categorical. Since CHAID is a<br />

non-parametric technique, no normalization of original variables is needed (Van Diepen <strong>and</strong><br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Extent <strong>and</strong> characteristics of mire habitats in Galicia (NW Iberian Peninsula)<br />

196<br />

Franses 2006). CHAID technique was applied using mire classes as response variables against<br />

the abovementioned collection of potential predictors. For the validation of the model, we<br />

r<strong>and</strong>omly split the original data in a training <strong>and</strong> validation with 50 % of the cases assigned to<br />

each subset. We finally generated a contingency matrix for the contrast of the observed <strong>and</strong><br />

predicted classes for the validation set, <strong>and</strong> subsequently we computed the KAPPA statistic<br />

(Bishop et al. 1975) as a measure of model accuracy.<br />

3. Results <strong>and</strong> Discussion<br />

Results of the CHAID classification are presented in figure 2. According the diagram, the most<br />

powerful discriminant variable was distance to sea. In the second level nodes other variables<br />

related to geology, topography <strong>and</strong> climate were taking into account for the discrimination of<br />

mire types. Finally, water balance was considered for the discrimination in the third level<br />

between some blanket bogs <strong>and</strong> Cladium fens with similar values regarding their distance to sea<br />

<strong>and</strong> slope. Overall accuracy reached more than 96 % (table 3) while the estimation of KAPPA<br />

index, regarded as a more reliable indication of the overall accuracy due to its compensation of<br />

chance agreement (Congalton 1991), achieved a value of 0.93, indicating an almost perfect<br />

agreement between the classification <strong>and</strong> reference values (L<strong>and</strong>is <strong>and</strong> Koch 1977). Per class<br />

accuracies reached particularly high values for blanket bogs <strong>and</strong> tufa, while fen <strong>and</strong> Cladium<br />

mires accuracies were lowered because the confusions with each other <strong>and</strong> also with tufa mires.<br />

According these results, blanket bogs occur on sub-coastal areas at different exposures, more<br />

frequently facing north <strong>and</strong> under high annual rainfall or water balance, as previously stated by<br />

biogeographic studies of this habitat in NW Iberian Peninsula (Rodríguez Guitián et al. 2007).<br />

Cladium fens are located close to the coastline, on sedimentary deposits (corresponding in most<br />

cases with the inl<strong>and</strong> border of coastal salt marshes). They also occur in the inl<strong>and</strong>, on flat<br />

surfaces under not particularly high water balances, where some confusion with tufa or fens<br />

could happen. Even when some fen localities may occur in the inl<strong>and</strong>, they tend to appear close<br />

to the coast, on ultra basic material sharing these environmental conditions with localities of<br />

Cladium fens <strong>and</strong> tufa formations. Finally, tufa occurs on the coast, in springs or water table<br />

ruptures leaching fossil/raised coastal dunes systems still rich in carbonates on coastal cliffs, or<br />

alternatively in the inl<strong>and</strong>, linked to the few ditches of limestone rocks in the region (Ramil<br />

Rego et al. 2008).<br />

4. Conclusions<br />

In the present work we explored the role of different environmental controls on the occurrence<br />

of four types of mire habitats. We found out that the degree of continentality (using the reliefcorrected<br />

distance to sea as an indicator) play a potential key role in the differences in spatial<br />

distribution of types in the region of Galicia. However mire types occurrence can not be<br />

differentiated or explained on the basis of just one kind of environmental control, but rather a<br />

combination of different controls (as geology, distance to sea, climate or lithology), being the<br />

importance of each one related to the ecology, tolerance <strong>and</strong> requirements of each particular<br />

habitat.<br />

The results corroborate in a quantitative <strong>and</strong> spatially explicit way the previous knowledge on<br />

the ecological differences between the different mire habitats in the region. This information has<br />

a potential application in habitat management plans for protected areas in combination to other<br />

datasets (e.g. spatial pattern <strong>and</strong> distribution, vulnerability <strong>and</strong> resilience, future climatic <strong>and</strong><br />

l<strong>and</strong> use scenarios) <strong>and</strong> also constitutes a first step towards a predictive biogeographic<br />

modelling of habitat distribution.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Extent <strong>and</strong> characteristics of mire habitats in Galicia (NW Iberian Peninsula)<br />

197<br />

References<br />

Biggs, D.; Deville, B. <strong>and</strong> Suen, E., 1991. A Method of Choosing Multiway Partitions for<br />

Classification <strong>and</strong> Decision Trees. Journal of Applied Statistics, 18: 49-62.<br />

Bishop, Y. M. M.; Fienberg, S. E. <strong>and</strong> Holl<strong>and</strong>, P. W., 1975. Discrete multivariate analysis :<br />

theory <strong>and</strong> practice. MIT Press, Cambridge. Massachusetts.<br />

Congalton, R. G., 1991. A review of assessing the accuracy of classifications of remotely sensed<br />

data. Remote Sensing of Environment, 37: 35-46.<br />

European Commission 2003. Interpretation Manual of European Union Habitats - EUR 25.<br />

European Commission. DG Environment. Nature <strong>and</strong> Biodiversity. Brussels. 129 pp.<br />

European Environment Agency, 2008. Biogeographical regions in Europe.<br />

http://www.eea.europa.eu/data-<strong>and</strong>-maps/figures/biogeographical-regions-in-europe (Access 03/01/2010)<br />

Goodwillie, R., 1980. Les Tourbières en Europe. Comite Europeen Pour la Sauvegarde de la<br />

Nature et des Ressources Naturelles. Conseil de L´Europe. Collection Sauvegarde de la<br />

Nature Nº 19. Strasbourg. 82 pp.<br />

Graniero, P. A. <strong>and</strong> Price, J. S., 1999. The importance of topographic factors on the distribution<br />

of bog <strong>and</strong> heath in a Newfoundl<strong>and</strong> blanket bog complex. CATENA, 36: 233-254.<br />

Izco Sevillano, J.; Díaz Varela, R. A.; Martínez Sánchez, S.; Rodríguez Guitián, M. A.; Ramil<br />

Rego, P. <strong>and</strong> Pardo Gamundi, I. M., 2001. Análisis y Valoración de la Sierra de O<br />

Xistral: un Modelo de Aplicación de la Directiva Hábitat en Galicia. Xunta de Galicia.<br />

Santiago de Compostela. 161 pp.<br />

Kass, G. V., 1980. An Exploratory Technique for Investigating Large Quantities of Categorical<br />

Data. Journal of Applied Statistics, 29: 119-127.<br />

L<strong>and</strong>is, J. R. <strong>and</strong> Koch, G. G., 1977. The measurement of observer agreement for categorical<br />

data. Biometrics, 33: 159-174.<br />

Martínez Cortizas, A. <strong>and</strong> Pérez Alberti, A. (Dir.), 1999. Atlas climático de Galicia. Xunta de<br />

Galicia. Santiago de Compostela. 207 pp.<br />

Raeymaekers, G., 2000. Conserving mires in the European Union. Actions Co-Financed by<br />

LIFE-Nature. European Commission. DG XI. 90 pp.<br />

Ramil-Rego, P. <strong>and</strong> Crecente Maseda, R. (Dir.), 2005. Planes de conservación de las ZEPVN de<br />

Galicia. Xunta de Galicia. Consellería de Medio Ambiente. Dirección Xeral de<br />

Conservación da Natureza. Santiago de Compostela.<br />

Ramil-Rego et al., 2008. Os hábitats de Interesse Comunitario en Galicia. Descrición e<br />

Valoración Territorial. Monografías do IBADER. Universidade de Santiago de<br />

Compostela. Lugo. Spain. 263 pp.<br />

Rivas-Martinez, S. <strong>and</strong> Rivas-Saenz, S., 2009. Worldwide Bioclimatic Classification System,<br />

1996-2009. Phytosociological Research Center, Spain. http://www.globalbioclimatics.org.<br />

Rodríguez Guitián, M.A; Ramil-Rego, P.; Real, C.; Díaz Varela, R.; Ferreiro da Costa, J. <strong>and</strong><br />

Cillero, C., 2009. Caracterización vegetacional de los complejos de turberas de cobertor<br />

activas del SW europeo. En: F. Llamas & C. Acedo (Coords.): Botánica Pirenaico-<br />

Cantábrica en el siglo XXI. Área de Publicaciones. Universidad de León. León: 633-653<br />

Rydin, H. <strong>and</strong> Jeglum, J., 2008. The biology of peatl<strong>and</strong>s. Oxford University Press. New York.<br />

343 pp.<br />

Rodríguez Guitián, M.A. <strong>and</strong> Ramil-Rego, P. 2008. Fitogeografía de Galicia NW Ibérico.<br />

análisis histórico y nueva propuesta corológica. Recursos Rurais, 14: 19-50.<br />

Van Diepen, M. <strong>and</strong> Franses, P.H., 2006. Evaluating chi-squared automatic interaction<br />

detection. Information Systems, 31: 814-831.<br />

Wainwright, J. <strong>and</strong> Mulligan, M. (Eds.), 2004. Environmental Modelling. Finding Simplicity in<br />

Complexity. John Willey <strong>and</strong> Sons, Ltd. West Sussex, Engl<strong>and</strong>. 408 pp.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Extent <strong>and</strong> characteristics of mire habitats in Galicia (NW Iberian Peninsula)<br />

198<br />

Table 1: Explanatory variables<br />

Variable group Variable Acronym Type Units<br />

Annual average temperature Temp Ordinal Adimensional (intervals)<br />

Climate<br />

Annual total rainfall Rain Ordinal Adimensional (intervals)<br />

Annual total evapotranspiration ETP Ordinal Adimensional (intervals)<br />

Annual water balance Waterbal Ordinal Adimensional (intervals)<br />

Transformed aspect TRASP Scale Adimensional<br />

Curvature Curv Scale Adimensional<br />

Plan curvature (across slope) Plancurv Scale Adimensional<br />

Topography<br />

Profile curvature (along slope) Profcurv Scale Adimensional<br />

Slope Slope Scale degrees<br />

Terrain shape index Tershp Scale Adimensional<br />

Wetness index Wenidn Scale Adimensional<br />

Elevation DEM Scale m a.s.l.<br />

Geology Geologic material MAT Nominal Adimensiona<br />

Distance to sea Distance to sea Distosea Scale m<br />

Table 2: Intervals for the climate variables<br />

Average year temperature Total year rainfall Total year Evapotranspiration Total year water balance<br />

Code Intervals (ºC) Code Intervals (mm) Code Intervals (mm) Code Intervals (mm)<br />

1 15 9 > 2000 - - - -<br />

Table 3. Accuracy of the CHAID classification<br />

Predicted<br />

Observed Blanket Cladium Fen Tuf Percent correct<br />

blanket 1767 0 0 1 99,9%<br />

cladium 13 122 4 40 68,2%<br />

fen 11 11 49 0 69,0%<br />

tuf 1 14 10 924 97,4%<br />

Percent correct 60,4% 5,0% 2,1% 32,5% 96,5%<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Extent <strong>and</strong> characteristics of mire habitats in Galicia (NW Iberian Peninsula)<br />

199<br />

Figure 1: Study area<br />

Figure 2: CHAID results<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Assessment of conservation status in managed chestnut forest<br />

200<br />

Assessment of conservation status in managed chestnut forest by<br />

means of l<strong>and</strong>scape metrics, physiographic parameters <strong>and</strong> textural<br />

features<br />

Ramón Alberto Diaz-Varela 1* , Pedro Álvarez-Álvarez 2 , Emilio Diaz-Varela 1 & Silvia<br />

Calvo-Iglesias 3<br />

1 University of Santiago de Compostela, Spain<br />

2 University of Oviedo, Spain<br />

3 University of Vigo, Spain<br />

Abstract<br />

In this work we aim to model the conservation status of chestnut forests in the NW of Iberian<br />

peninsula, assuming the coverage of chestnut <strong>and</strong> developing state as indicator of forest st<strong>and</strong>s<br />

quality. We used as potential predictors features available at wide geographic scales, namely:<br />

geographic location, l<strong>and</strong>-cover heterogeneity approached by means of satellite images texture,<br />

<strong>and</strong> l<strong>and</strong>scape metrics. As some of these variables are prone to be influenced by terrain shape,<br />

we also introduced topographic parameters derived from Digital Elevation Models in the<br />

analyses. Results allowed us to classify chestnut forests in relation to their conservation status<br />

<strong>and</strong> to identify current degradation trends. We expect that the results would serve as a basis for a<br />

more indeep research on conservation strategies for Iberian chestnut forests.<br />

Keywords: Chestnut; NW Iberian Peninsula; L<strong>and</strong>scape metrics; Terrain features; Image<br />

texture<br />

1. Introduction<br />

Chestnut forest has been recognised as habitat of interest in the European Natura 2000 network<br />

<strong>and</strong> it was also acknowledged as a characteristic cultural l<strong>and</strong>scape of the Mediterranean <strong>and</strong><br />

Atlantic regions in Europe. Despite of its environmental <strong>and</strong> cultural interest it is threatened by<br />

pathogen fungi, traditional management ab<strong>and</strong>onment <strong>and</strong> recent rural l<strong>and</strong>scape change both in<br />

the Atlantic <strong>and</strong> Mediterranean Europe (Cevasco, 2009; Díaz Varela et al. 2009). In fact, its<br />

conservation status is regarded as “not favourable” either in the Alpine, Continental <strong>and</strong><br />

Mediterranean regions, while information is lacking in the Atlantic region, according to the<br />

European Topic Centre on Biological Diversity (2008). One of the main threats for the long<br />

term conservation of this habitat is the natural evolution of the chestnut st<strong>and</strong>s towards some<br />

kind of native forest due to natural succession dynamics or its invasion by alien woodl<strong>and</strong><br />

species.<br />

Taking these threats into account, it is necessary to design simple, repeatable <strong>and</strong> precise<br />

methods for the monitoring <strong>and</strong> assessment of chestnut forest conservation status at large scale<br />

in the European Union in general <strong>and</strong> in the Atlantic region in particular. In this work we<br />

explore the possibilities of automatic <strong>and</strong> spatially exhaustive assessment of chestnut woodl<strong>and</strong>s<br />

conservation status using datasets easily available like remote sensed imagery, forest inventory<br />

maps <strong>and</strong> digital terrain models. We hypothesized that some simple st<strong>and</strong> variables are reliable<br />

* Corresponding author. Tel.: +34 982 285 900 ext. 22482 - Fax: +34 982 285 985<br />

Email address: ramon.diaz@usc.es<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Assessment of conservation status in managed chestnut forest<br />

201<br />

indicators of the conservation status of chestnut woodl<strong>and</strong>s. Then using statistic methods, we<br />

explored the predictor value of several datasets potentially related to forest structure <strong>and</strong><br />

composition, like patch morphology metrics (Saura <strong>and</strong> Carballal 2004) or satellite image<br />

texture (Kayitakire et al. 2006). As these potential predictor features are prone to be influenced<br />

by topography, we also considered terrain attributes in the analysis.<br />

We conducted the study in a sector of NW Iberian Peninsula, an area where good examples of<br />

chestnut forest st<strong>and</strong>s with different conservation status occur. In order to provide a dataset with<br />

an adequate size to assemble variability in forest st<strong>and</strong> conditions, we considered an area of<br />

approximately 24 448 km 2 corresponding with the l<strong>and</strong> extent of the L<strong>and</strong>sat TM scene 204-30<br />

(Fig. 1). Altitudes range from sea level to about 2000 m a.s.l. showing a contrasted relief.<br />

Biogeographically most of the area is included in the Atlantic Region, but a sector of the SE<br />

quadrant of the scene belongs to the Mediterranean region, (European Environmental Agency<br />

2008; Rivas-Martinez <strong>and</strong> Rivas-Saenz, 2009). Natural vegetation is dominated by different<br />

deciduous forest, mostly dominated by Quercus robur <strong>and</strong> Quercus petraea, coexisting toward<br />

the west with Betula alba <strong>and</strong> Fagus sylvatica <strong>and</strong> interspersed with mixed mesophytic<br />

woodl<strong>and</strong> at lower altitudes. Transition to Continental <strong>and</strong> Mediterranean environments are<br />

characterised by Quercus pyrenaica forests, while some evergreen or sclerophyllous forest are<br />

restricted to the dryer <strong>and</strong> warmer locations (Rivas-Martínez, 2007). Chestnut forests appear as<br />

pure st<strong>and</strong>s or interspersed with different deciduous forests or alien species in the area, being<br />

more frequent towards the East.<br />

2. Methodology<br />

Chestnut forest variables were extracted from the Digital <strong>Forest</strong> Map of Spain at a scale 1:50<br />

000 (Ministerio de Medio Ambiente, 2002). This map is based on digitalisation of forests from<br />

satellite images or aerial orthophotographs <strong>and</strong> supported by fieldwork <strong>and</strong> ancillary data.<br />

Minimum mapable area is 2.5 has for forested area <strong>and</strong> 6.25 has for other l<strong>and</strong> cover types. The<br />

map provides information on tree species composition, overall (in percentage) <strong>and</strong> per-species<br />

(in 10% intervals) canopy coverage, along with other st<strong>and</strong> variables. We queried the map to<br />

extract the chestnut forest patches for the study area. At the effects of the present work <strong>and</strong> as<br />

we are focused on dense forest with a significant share of chestnut in the canopy, we considered<br />

patches with minimum crown coverage of tree species of 60 %, a minimum of 20 % of chestnut<br />

coverage, <strong>and</strong> the interval 2 to 4 of stage of st<strong>and</strong> development for chestnut out of a scale from 1<br />

to 4, resulting a total of 1585 chestnut patches. For each patch we retrieved chestnut canopy<br />

coverage in intervals of 10%, coded as a scale variable from 1 (0-10% to 10 (90-100%) along<br />

with the stage of st<strong>and</strong> development. Then we set three classes of chestnut st<strong>and</strong> quality based<br />

on the combination of these two variables (cf. table 1), being class 1 the worst <strong>and</strong> 3 the best<br />

quality. As some of the analysis inputs must be in raster format, we also generated 30 m<br />

resolution raster mask of chestnut patches.<br />

A collection of morphometric parameters were extracted from the selected forest patches by<br />

means of two different approaches (table 2). The first approach was based on the use six<br />

different l<strong>and</strong>scape metrics: patch area (MPS), patch edge (TE), shape index (MSI), perimeter<br />

area ratio (MPAR), fractal dimension (MFRACT), <strong>and</strong> number of shape characteristic points<br />

(NSCP). All were calculated using the software V-Late (Lang <strong>and</strong> Tiede 2003), except for<br />

NSCP (Moser et al. 2002). Calculation was made at patch level. We also used the GUIDOS<br />

software (Graphical User Interface for the Description of image Objects <strong>and</strong> their Shapes)<br />

initially designed for morphological spatial pattern analysis (MSPA) of forest functional<br />

connectivity (Vogt et al., 2009) as an alternative morphometic approach. The output of the<br />

MSPA is a raster layer where patch cells are assigned to seven morphological spatial pattern<br />

classes: edge, core, perforated, islets, bridge, loop <strong>and</strong> branch. We ran the software using the<br />

default parameters for the computation of MSPA, considering one pixel edge (30 m).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Assessment of conservation status in managed chestnut forest<br />

202<br />

Considering that terrain relief might play a key role in the l<strong>and</strong>scape structure <strong>and</strong> in vegetation<br />

pattern in general (Hoechstetter, et al. 2008) we selected several terrain features that showed<br />

their potential as predictors for forest distribution in previous works (Lindenmayer et al. 1999).<br />

Terrain features were derived from a raster digital elevation model with a spatial resolution of<br />

30 m. We selected a number of first <strong>and</strong> second order terrain features widely used in<br />

hydrological, geomorphological, <strong>and</strong> ecological studies (Wilson <strong>and</strong> Gallant 2000) along with<br />

other addressing more specifically vegetation <strong>and</strong> forest assessment (Mcnab 1989; Roberts <strong>and</strong><br />

Cooper 1989) as detailed in table 2.<br />

Image texture has shown its potential as predictor for forest st<strong>and</strong> characteristics (see p.e.<br />

Franklin et al., 2001). In order to assess its value as predictor of chestnut st<strong>and</strong> quality at patch<br />

level, we computed eight texture features based on kernel grey level co-occurrence matrices as<br />

proposed by Haralick et al., (1973) on a NDVI (greenness index) image, using the software<br />

package PCI TM 9.1 (cf. table 2) setting a direction 45º, displacement of one pixel <strong>and</strong> a window<br />

size of 5x5. We used a relatively small size for the kernel as some small or narrow patches are<br />

foreseen <strong>and</strong> to avoid influence from the neighbouring patches.<br />

As terrain relief <strong>and</strong> texture information are refered at pixel level, we computed mean (MN) <strong>and</strong><br />

st<strong>and</strong>ard deviation (SD) of these features for each patch in order to obtain information at patch<br />

level. The final dataset comprises 16 texture variables (MN <strong>and</strong> SD for the eight texture<br />

features), 14 terrain variables (MN <strong>and</strong> SD for the seven terrain features) <strong>and</strong> 13 patch<br />

morphology variables (six patch l<strong>and</strong>scape metrics along with the percentages of each of the<br />

seven typologies of the MPSA analysis for a given forest patch). These 43 variables were<br />

analysed by means of a Classification <strong>and</strong> Regression Tree (CaRT) procedure using the<br />

statistical package SPSS TM 16.0. CaRT is a data mining technique pursuing the recursive binary<br />

partition of a multivariate space of independent or predictor variables into regions for which the<br />

values of the dependent or response (in this case st<strong>and</strong> quality) variable are approximately equal.<br />

In this application, we use the method for exploring the internal structure of the dataset in order<br />

to assess the potential of the different variables as chestnut patch quality predictors.<br />

3. Results<br />

Figure 2 shows the results of the CaRT application on the 43 potential predictors, pruned to 3<br />

branch levels. The first significant classification variable was mean terrain slope, with a value of<br />

15º that splits a terminal node corresponding with patches of gently slope assigned mainly<br />

(72 %) to the worst class of chestnut st<strong>and</strong>s quality. The second significant classification<br />

variable was the mean patch value of GLCM mean (Mean_MN), a variable sensitive to the<br />

image tone <strong>and</strong> texture, that could be interpreted in this case as an indicator of the degree of<br />

texture complexity <strong>and</strong> greenness inside a patch. Values lower that 48 for this variable led to a<br />

third level defined by terrain elevation, spitting at a value of approximately 600 m in two<br />

branches, one dominated by the lower quality classes corresponding to low altitudes <strong>and</strong> other<br />

dominated by high quality st<strong>and</strong>s, located at higher altitudes. Values of Mean_MN higher that<br />

48 led to another splitting level, where the criterion was patch size. In this case low area patches<br />

correspond with good quality st<strong>and</strong>s, while larger patches correspond to st<strong>and</strong>s containing lower<br />

chestnut canopy coverage <strong>and</strong> development (qualities 2 <strong>and</strong> 3).<br />

4. Conclusions<br />

We assessed the potential discrimination power of different kind of variables <strong>and</strong> chestnut forest<br />

conservation status by means of CaRT data mining. Despite of the relative high degree of<br />

impurity in some of the nodes, the method allowed us to recognise some mayor trends of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Assessment of conservation status in managed chestnut forest<br />

203<br />

behaviour of predictors <strong>and</strong> chestnut forest quality classes. Focusing on high quality st<strong>and</strong>s, they<br />

occur mainly on quite steep slopes <strong>and</strong> also tend to occur at altitudes higher than 600 m showing<br />

a relative low textured <strong>and</strong> low greenness value. This is consistent with the theoretical<br />

confinement of the chestnut traditional forest to mountain or marginal areas, where agricultural<br />

activity is constrained by the environment, <strong>and</strong> also indicates that a relative low degree of<br />

heterogeneity in the patches might be used as an indicator of well preserved st<strong>and</strong>s. Good<br />

quality st<strong>and</strong>s can also occur on more textured (heterogeneous) patches, but in this case<br />

corresponding to small plots, fact that could be interpreted as the separation between large<br />

heterogeneous forest patches with lower relative chestnut coverage <strong>and</strong> small plots dominated<br />

by chestnut but with some heterogeneity in the st<strong>and</strong> characteristics.<br />

References<br />

Cevasco, R., 2009. Terraced Castanea woods in Val di Vara, Liguria, NW Italy. In Krzywinski,<br />

K.; O'connell, M. <strong>and</strong> Küster, H., Eds. Cultural <strong>L<strong>and</strong>scapes</strong> of Europe. Fields of Demeter<br />

Haunts of Pan. Aschembeck Media UG. Bremen, Pp: 106-107.<br />

Diaz Varela, R. A.; Calvo Iglesias, M. S.; Diaz Varela, E. R.; Ramil Rego, P. <strong>and</strong> Crecente<br />

Maseda, R., 2009. Castanea sativa forests: a threatened cultural l<strong>and</strong>scape in Galicia, NW<br />

Spain. In Krzywinski, K.; O'connell, M. <strong>and</strong> Küster, H., Eds. Cultural <strong>L<strong>and</strong>scapes</strong> of<br />

Europe. Fields of Demeter Haunts of Pan. Aschembeck Media UG. Bremen. Pp: 106-107.<br />

European Environment Agency, 2008. EEA Biogeographical regions of Europe 2008.<br />

http://www.eea.europa.eu/data-<strong>and</strong>-maps/data/biogeographical-regions-europe-2008.<br />

European Topic Centre on Biological Diversity, 2008. Article 17 Technical Report, 2001-2006).<br />

European Commission .DG Environment. http://biodiversity.eionet.europa.eu/article17<br />

Franklin, S. E.; Wulder, M. A. <strong>and</strong> Gerylo, G. R., 2001. Texture analysis of IKONOS<br />

panchromatic data for Douglas-fir forest age class separability in British Columbia.<br />

International Journal of Remote Sensing, 22: 2627 - 2632.<br />

Haralick, R.; Shanmugam, K. <strong>and</strong> Dinstein, I., 1973. Textural features for image classification.<br />

IEEE Transactions on Systems, Man, <strong>and</strong> Cybernetics, 3: 610-621.<br />

Hoechstetter, S.; Walz, U.; Dang, L. H. <strong>and</strong> Thinh, N. X., 2008. Effects of topography <strong>and</strong><br />

surface roughness in analyses of l<strong>and</strong>scape structure – A proposal to modify the existing<br />

set of l<strong>and</strong>scape metrics. L<strong>and</strong>scape Online, 3: 1-14.<br />

Kayitakire, F.; Hamel, C. <strong>and</strong> Defourny, P., 2006. Retrieving forest structure variables based on<br />

image texture analysis <strong>and</strong> IKONOS-2 imagery. Remote Sensing of Environment, 102:<br />

390-401.<br />

Lang, S.; Tiede, D. (2003): vLATE Extension für ArcGIS - vektorbasiertes Tool zur<br />

quantitativen L<strong>and</strong>schaftsstrukturanalyse, ESRI Anwenderkonferenz 2003 Innsbruck.<br />

CDROM.<br />

Lindenmayer, D. B.; Mackey, B. G.; Mullen, I. C.; Mccarthy, M. A.; Gill, A. M.; Cunningham,<br />

R. B. <strong>and</strong> Donnelly, C. F., 1999. Factors affecting st<strong>and</strong> structure in forests - are there<br />

climatic <strong>and</strong> topographic determinants. <strong>Forest</strong> Ecology <strong>and</strong> Management, 123: 55-63.<br />

Mcnab, W. H., 1989. Terrain Shape Index: Quantifying Effect of Minor L<strong>and</strong>forms on Tree<br />

Height. <strong>Forest</strong> Science, 35: 91-104.<br />

Moser, D.; Zechmeister, H.G.; Plutzar, C.; Sauberer, N.; Wrbka, T.; Grabherr, G. (2002)<br />

<strong>L<strong>and</strong>scapes</strong> patch shape complexity as an effective measure for plant species richness in<br />

rural l<strong>and</strong>scapes. L<strong>and</strong>scape Ecology 17:657–669<br />

Ministerio De Medio Ambiente, 2002. Mapa <strong>Forest</strong>al de España escala 1:50 000. Banco de<br />

Datos de la Naturaleza. Organismo Autónomo de Parques Naturales. Ministerio de Medio<br />

Ambiente. Madrid.<br />

Rivas-Martinez, S. <strong>and</strong> Rivas-Saenz, S., 2009. Worldwide Bioclimatic Classification System,<br />

1996-2009. Phytosociological Research Center, Spain. http://www.globalbioclimatics.org.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Assessment of conservation status in managed chestnut forest<br />

204<br />

Rivas-Martínez, S., 2007. Mapa de series, geoseries y geopermaseries de vegetación de España.<br />

Memoria del mapa de vegetación potencial de España. Itinera Geobotanica, 17: 5-436.<br />

Roberts, D. W. <strong>and</strong> Cooper, S. V., 1989. Concepts <strong>and</strong> techniques of vegetation mapping. L<strong>and</strong><br />

Classifications Based on Vegetation. Applications for Resource Management. USDA<br />

<strong>Forest</strong> Service GTR INT-257. Ogden, UT. Pp: 90-96.<br />

Saura, S. <strong>and</strong> Carballal, P, 2004. Discrimination of native <strong>and</strong> exotic forest patterns through<br />

shape irregularity indices: An analysis in the l<strong>and</strong>scapes of Galicia, Spain. L<strong>and</strong>scape<br />

Ecology. 19: 647-662.<br />

Vogt, P.; Ferrari, J. R.; Lookingbill, T. R.; Gardner, R. H.; Riitters, K. H. <strong>and</strong> Ostapowicz, K.,<br />

2009. Mapping functional connectivity. Ecological Indicators, 9: 64-71.<br />

Wilson, J. P. <strong>and</strong> Gallant, J. C., 2000. Digital Terrain Analysis. In Wilson, J.P. <strong>and</strong> Gallant, J.C.,<br />

Eds: Terrain Analysis: Principles <strong>and</strong> Applications. John Wiley <strong>and</strong> Sons, INC. New<br />

York. Pp: 1-28.<br />

Table 1: Criteria for st<strong>and</strong> quality index. Grey tone indicates canopy coverage <strong>and</strong> stage of st<strong>and</strong><br />

development corresponding to each of the chestnut forest quality classes: light grey quality 1;<br />

medium grey quality 2 <strong>and</strong> dark grey quality 3. Cell values indicates number of patches.<br />

Chestnut canopy coverage<br />

2 3 4 5 6 7 8 9 10<br />

Stage of 2 - 5 2 1 1 1 - - -<br />

st<strong>and</strong> 3 1 56 33 31 40 17 36 5 1<br />

development<br />

4 1 345 202 174 149 152 131 151 50<br />

Table 2: Predictors<br />

Variable group Variable Acronym Reference<br />

patch area (m 2 ) MPS Lang <strong>and</strong> Tiede 2003<br />

patch edge (m) TE Lang <strong>and</strong> Tiede 2003<br />

shape index MSI Lang <strong>and</strong> Tiede 2003<br />

perimeter area ratio MPAR Lang <strong>and</strong> Tiede 2003<br />

fractal dimension MFRACT Lang <strong>and</strong> Tiede 2003<br />

L<strong>and</strong>scape pattern<br />

number of shape characteristic points NSCP Moser et al. 2002<br />

Fraction of Edge Edge Vogt et al., 2007, 2009<br />

Fraction of Core Core Vogt et al., 2007, 2009<br />

Fraction of Perforated Perforated Vogt et al., 2007, 2009<br />

Fraction of Islets Islets Vogt et al., 2007, 2009<br />

Fraction of Bridge Bridge Vogt et al., 2007, 2009<br />

Fraction of Loop Loop Vogt et al., 2007, 2009<br />

Fraction of Branch Branch Vogt et al., 2007, 2009<br />

Elevation (m a.s.l.) DEM Wilson <strong>and</strong> Gallant 2000<br />

Curvature Curv Wilson <strong>and</strong> Gallant 2000<br />

Plan curvature Plancurv Wilson <strong>and</strong> Gallant 2000<br />

Terrain<br />

Profile curvature Profcurv Wilson <strong>and</strong> Gallant 2000<br />

Slope (degrees) Slope Wilson <strong>and</strong> Gallant 2000<br />

Transformed aspect TRASP Roberts <strong>and</strong> Cooper 1989<br />

Terrain shape index Tershp McNab 1989<br />

Contrast Contr Harlick 1973<br />

Correlation Corr Harlick 1973<br />

Dissimilarity Dissim Harlick 1973<br />

Image texture<br />

Entropy Entrop Harlick 1973<br />

Homogeneity Homog Harlick 1973<br />

Mean Mean Harlick 1973<br />

Second Moment Sec_mom Harlick 1973<br />

Variance St<strong>and</strong>_de Harlick 1973<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Assessment of conservation status in managed chestnut forest<br />

205<br />

Chestnut Woodl<strong>and</strong><br />

Figure 1: Study area<br />

Figure 2: CART results<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.A.E. Diogo & J. Aranha 2010. GIS analysis of the Antidote Programme in Portugal<br />

206<br />

GIS analysis of the Antidote Programme in Portugal<br />

Patrícia A.E. Diogo 1 & José Aranha 1,2*<br />

1 Departamento de Ciências Florestais e Arquitectura Paisagista, Universidade de Trásos-Montes<br />

e Alto Douro, 5001-801 Vila Real, Portugal<br />

2 CITAB, Universidade de Trás-os-Montes e Alto Douro, 5001-801 Vila Real, Portugal<br />

Abstract<br />

The aim of this research was to analyse the huge number of animal poisoning occurrences in<br />

Portugal <strong>and</strong> to establish a relationship between poisoning, l<strong>and</strong> use, l<strong>and</strong> cover <strong>and</strong> human<br />

activity.<br />

It was used a large database, belonging to the programme Antidote Portugal in order to create a<br />

GIS. This database records all occurrences attributes, such as, the location of dead animals, the<br />

number of affected individuals, the species that belong <strong>and</strong> their location.<br />

Two approaches were made, one based on poisoning occurrences <strong>and</strong> one base on number of<br />

dead animal per occurrence, in order to calculate hazard maps. Both approaches were based on<br />

multivariate analysis <strong>and</strong> geostatistical processes.<br />

According to the results, agro-forested areas are those were the most cases occur. A special<br />

reference must be made to the Protected Areas <strong>and</strong> the Natural Park of Southwest Alentejo <strong>and</strong><br />

Costa Vicentina, were token place few cases but very deadly.<br />

Keywords: Principal Component Analysis, Geostatistic, Hazard Poisoning, Programme<br />

Antidote Portugal, GIS<br />

1. Introduction<br />

The Portuguese Antidote Programme (PAP) was created in January the 12th, 2003 with the aim<br />

of evaluating the effects of the use of poisons on wildlife populations <strong>and</strong> to establish measures<br />

to control this problem (Br<strong>and</strong>ão 2003). It is a platform based on the Spanish Antidote<br />

Programme (SAP), which began in 1997 due to growing concern about the illegal use of<br />

poisons <strong>and</strong> other toxic substances posing a threat to wildlife conservation (FCQ 2009). The<br />

PAP’s aims is to establish a cooperation protocol between different organizations, public <strong>and</strong><br />

private entities, <strong>and</strong> to developed policies in order to gather all the scattered information <strong>and</strong><br />

establish mechanisms for its concentration <strong>and</strong> a complete study (Br<strong>and</strong>ao 2003).<br />

The first use of poison is reported to the nineteenth century, leading some drastic declines or<br />

extinction of some wildlife populations, especially the Iberian Wolf (Canis lupus) <strong>and</strong> carrion<br />

birds. In agro forested areas <strong>and</strong> natural areas, poisons area used for various reasons, mainly as<br />

an attempt to control predators of game animals <strong>and</strong> livestock. These actions are carried out by<br />

hunters <strong>and</strong> hunting areas managers, <strong>and</strong> have as target species <strong>and</strong> feral dogs, wolves <strong>and</strong><br />

mammals of small <strong>and</strong> medium businesses. This is the most accessible <strong>and</strong> successful method<br />

because of the easiness of which it can be applied <strong>and</strong> the number of individuals that can be<br />

eliminated under a low level of effort for anyone who applies. Almost all historical references<br />

* Corresponding author; Telf. + 00 351 259 350 856 - Fax. + 00 351 250 350 480<br />

Email address: j.aranha.utad@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.A.E. Diogo & J. Aranha 2010. GIS analysis of the Antidote Programme in Portugal<br />

207<br />

refer the use of strychnine. This poison is often used in horse carcasses, viscera of ruminants<br />

<strong>and</strong> dogs, <strong>and</strong> chicken gizzards, thus attracting not only species such as wolves, but also carrion<br />

birds (Br<strong>and</strong>ão, 2004).<br />

National Association for Nature Conservation (Quercus - Associação Nacional de Conservação<br />

da Natureza) is the organisation which supports the project, as been developing a database of all<br />

occurrences recorded since 1992 until now. This database records: place where the animal was<br />

found, who found it, the species, the cause of death, etc. This important database has hundreds<br />

of records, <strong>and</strong> an enormous potential for a Geographical Information System (GIS) creation<br />

(Diogo 2009).<br />

The original aim, of the present research, was to create a GIS, which ultimately aim was to<br />

produce maps of the main risk of poisoning for the entire Portuguese country (see Figure 1,<br />

Portuguese isl<strong>and</strong> not included).<br />

To achieve those aims, it was necessary to stablish a relationship between the poisoning<br />

episodes, environmental characteristics <strong>and</strong> the human activity. Thus, they were used GIS<br />

techniques, such as 3D Analyst, Spatial Analyst <strong>and</strong> Geostatistical Analysis were used in ity<br />

(Morrison, 1991; Reis, 1997; Hoef et al, 2001; Clemente et al. 2002; Blanquer et al. 2005;<br />

Soares 2006; Abdi & N<strong>and</strong>ipati 2009).<br />

Figure 1: Study area location [Adapted from Portuguese Environment Atlas 2009]<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.A.E. Diogo & J. Aranha 2010. GIS analysis of the Antidote Programme in Portugal<br />

208<br />

2. Methodology<br />

Using the Quercus database for Portuguese Antidote Programme (PAP), all recorded<br />

information was used in order to create a geographic information system (GIS).<br />

In a second stage, the GIS was updated with information concerning:<br />

• L<strong>and</strong> use <strong>and</strong> l<strong>and</strong> cover map<br />

• Settlements Location<br />

• Portuguese Roads network<br />

• Digital Elevation Model<br />

In the third stage, 3D Analyst <strong>and</strong> Spatial Analyst Tools, were used in order to process<br />

information <strong>and</strong> to calculate:<br />

• L<strong>and</strong> slope <strong>and</strong> aspect<br />

• Distances to water bodies<br />

• Distance to road network<br />

• Distance to settlements<br />

Then, Principal Components Analysis, Cluster Analisys <strong>and</strong> Geostatistical Analysis were used<br />

in order to stablish a relationship between the poisoning episodes, environmental characteristics<br />

<strong>and</strong> the human activity (Morrison, 1991; Reis, 1997; Hoef et al, 2001; Soares 2006).<br />

In a fourth stage, previous present data were submited to geostatistical calculation, in order to<br />

create poisoning hazard maps for the entire Portuguese country (isl<strong>and</strong>s not included).<br />

3. Results<br />

The results showed a greater number of episodes (274) <strong>and</strong> poisoned individuals (781) were<br />

located in Agriculture Areas, which corresponds to an average of 2.85 dead animals per<br />

poisoning episode. The Artificial Territories (e.g. industrial areas) <strong>and</strong> <strong>Forest</strong>ed <strong>and</strong> Seminatural<br />

Areas presented similar values for the number of dead animals (331 <strong>and</strong> 371<br />

respectively), but significant differences for the number of episodes, 104 over 79 respectively.<br />

Thus, the rate of poisoned individuals per poisoning episode is higher for <strong>Forest</strong>ed <strong>and</strong> Seminatural<br />

Areas (4.70) then for Artificial Territories (3.18). Classes "Wetl<strong>and</strong>s" <strong>and</strong> "Water<br />

Bodies" do not present any type of episode.<br />

These results enable to state that the Agriculture Areas are those with higher probability to<br />

observe poisoning episodes but under small ratio of dead animals per episode (2.85). In the<br />

opposite, <strong>Forest</strong>ed <strong>and</strong> Semi-natural Areas are those with lower probability to observe a<br />

poisoning episode but under high ratio of dead animals per episode (4.70).<br />

4. Discussion<br />

Wild species.<br />

In "Artificialized Areas”, the most affected wild species are the Miliaria cal<strong>and</strong>ra with 6<br />

poisoned animals (16.67%), followed by Gyps fulvus <strong>and</strong> Streptopelia turtur both with 5<br />

poisoned animals (13.89% each) <strong>and</strong> the Ciconia ciconia with 4 dead animals (11.11%). The<br />

remaining 12 cases were attributed to other 12 species.<br />

In “Agricultural Areas”, the Gyps fulvus was the species that accounts for a greater number of<br />

poisoned animals (69) with nearly 44.52% of the cases, followed by the Canis lupus with 17<br />

cases (10.97%) <strong>and</strong> Ciconia ciconia 16 cases. Although the Milvus milvus presented a smaller<br />

number of poisoned animals (14 - 9.03%), this is quite worrying, especially when considering<br />

that Portugal only has 50 to 100 breeding pairs (Cabral et al. 2006).<br />

In class "<strong>Forest</strong>ed <strong>and</strong> Semi-natural Areas", Canis lupus was the species most affected, with<br />

29.73% of the deaths, followed by Aegypius monachus with 6 poisoned animals (16.22%) <strong>and</strong><br />

the species Gyps fulvus <strong>and</strong> Buteo buteo both with 5 dead animals (13.51% each). The<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.A.E. Diogo & J. Aranha 2010. GIS analysis of the Antidote Programme in Portugal<br />

209<br />

remaining 4 episodes of poisoning have been attributed to Aquila chrysaetos, Ciconia ciconia,<br />

Milvus milvus, <strong>and</strong> Milvus migrans.<br />

Game hunting species.<br />

In the class "Artificial Areas", Vulpes vulpes was the species clearly more affected in poisoning<br />

episodes, with 14 dead animals (87.50%), followed by Pica pica <strong>and</strong> Herpestes ichneumon with<br />

6.25% of the cases each.<br />

In class "Agricultural Areas," the Vulpes vulpes was, again, the species with the highest number<br />

of dead animals (71 or 85.54% of total), followed by Corvus corone <strong>and</strong> Pica pica with 6 <strong>and</strong> 4<br />

dead animals, respectively.<br />

In class "<strong>Forest</strong>ed <strong>and</strong> Semi-natural Areas", the Vulpes vulpes remains the species most affected<br />

(21 to 85.54%) followed by Herpestes ichneumon with 1 dead animal.<br />

Home Species.<br />

In "Artificial Areas", the Canis lupus familiaris was the species most affected with 234<br />

(83.87%) poisoned animals, followed by Columba livea with 9.32% (26 individuals) <strong>and</strong> by<br />

Felis silvestris catus with 19 dead animals (6.81%) <strong>and</strong><br />

In the "Agricultural Areas" the Canis lupus familiaris was, again, the species with the highest<br />

number of dead animals (495 - 90%). The Columba livea was the second most affected species<br />

(25 individuals), followed by the Felis silvestris catus with 18 cases. There is also the death of 4<br />

Ovis aries, which corresponds to a percentage lower than 1% of the total.<br />

In “<strong>Forest</strong>ed <strong>and</strong> Semi-natural Areas", only Canis lupus familiaris (308 to 98.72%) <strong>and</strong> Felis<br />

silvestris catus (4) dead animals were noticed.<br />

References<br />

Almeida, R., Bastos, W. R., Bernardi, J. V. E., Carvalho, D. P., Nascimento, E. L., Oliveira, R.<br />

C. 2007. Método geoestatístico para a modelagem ambiental de poluentes em sistemas<br />

lacustres – Amazónia Ocidental. Universidade Federal de Rondônia , Laboratório de<br />

Biogeoquímica Ambiental - Campus José ribeiro Filho, Porto Velho. Brasil.<br />

Abdi, A, e N<strong>and</strong>ipati, A. 2009. Bird diversity modeling using Geostatistics anf GIS. 12th<br />

AGILE International Conference on Geographic Information Science 2009. Leibniz<br />

Universitat Hannover, Germany.<br />

Bernhardsen, T. 1999. Geographic Information Systems, An Introduction – Second Edition.<br />

John Wiley & Sons, Inc. New York. U.S.A.<br />

Blanquer, J., Mateu, A., Cassiraga, E. <strong>and</strong> Romero, F. 2005. Generation of risk maps of<br />

contaminated areas combining Geographic Information Systems with Geostatistics.<br />

Universidad Politécnica de Valencia, Spain.<br />

Br<strong>and</strong>ão, R. 2003. Mortalidade de fauna por envenenamento de 1992 – 2002. Programa<br />

Antídoto - Portugal: I Relatório Técnico. Vila Real. Portugal.<br />

Br<strong>and</strong>ão, R. 2004. Estratégia nacional contra o uso de venenos. Programa Antídoto - Portugal.<br />

Vila Real. Portugal.<br />

Brio, R. G. del 1984. El lobo ibérico. Biología y mitología. Série Ciências de la Naturaleza, ed.<br />

Hermann Blume, Madrid, 344 pp.<br />

Burrough, P. A. 1996. Principles of geographical information systems for l<strong>and</strong> resources<br />

assessment. Oxford: Clarendon Press. Oxford University. New York.<br />

Clemente, R., Vieira, S., Reynolds, W., Jong, R. <strong>and</strong> Toop, G. 2002. Mapping groundwater<br />

pollution risk within an Agricultural watershed using Modeling, Geostatistics <strong>and</strong> GIS.<br />

Intituto Agronomico, São Paulo, Brasil.<br />

Fundación para la Conservación del Quebrantahuesos 2009. Proyectos – Justificación del<br />

Proyecto Programa Antídoto. http://www.quebrantahuesos.org/htm/es/otros/antidoto.htm<br />

read at 15/05/2009.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.A.E. Diogo & J. Aranha 2010. GIS analysis of the Antidote Programme in Portugal<br />

210<br />

Goovaerts, P. 1997. Geoestatistics for Natural Resources Evaluation. Oxford University Press.<br />

New York. U.S.A.<br />

Hengl, T. 2007. A Practical guide to Geoestatistical Mapping of Environmental Variales.<br />

European Commission, Joint Research Center – Institute for Environment <strong>and</strong><br />

Sustainability. Office for Official Publications oh the European Communities.<br />

Luxemgourg.<br />

Jakob, A. A. E., Young, A. F. 2006. O uso de métodos de interpolação espacial de dados nas<br />

análises sociodemográficas. Trabalho apresentado no XV Encontro Nacional de Estudos<br />

Populacionais, ABEP, realizado em Caxambu – MG – Brasil, de 18 a 22 de Setembro de<br />

2006.<br />

Júnior, S. R. (1934) – Aves de Portugal. XV Accipitriformes. Araújo & Sobrinho. Porto.<br />

Knotters, M., Brus, D. J., Voshaar, J. H. O. 1995. A comparison of kriging, co-kriging <strong>and</strong><br />

kriging combined with regression for spatial interpolation of horizon depth with censored<br />

observations. Geoderma. Elsevier, Amsterdam.<br />

Liverman, D.,Moran,E.F.,Rindfuss, R. R.,Stern, P.C. 1998. People <strong>and</strong> Pixels. National<br />

Academy Press. Washington, D.C. U.S.A.<br />

Matos, J. L. 2008. Fundamentos de Informação Geográfica – 5º edição actualizada e<br />

aumentada. LIDEL. Lisboa. Portugal.<br />

O’Sullivan, D. e Unwin, D. 2003. Geographic Information Analysis. John Wiley & Sons, Inc.<br />

New York. U.S.A.<br />

PAP (Programa Antídoto Portugal), 2003. I Relatório Técnico – Mortalidade de fauna por<br />

envenenamento 1992-2002, Março de 2003. Vila Real. Portugal.<br />

PAP (Programa Antídoto Portugal), 2008. Relatório de actividades e resultados, Dezembro de<br />

2008. Gouveia. Portugal.<br />

Salah, H. 2009. Geotatistical analysis of groundwater levels in the south Al Jabal Al Akhdar<br />

area using GIS. Water resource department, General Water Athority, Libya.<br />

Silva, L. G. C. 2007. Zonamento do risco de ocorrência do mal das folhas da Seringueira com<br />

base em sistemas de informação geográfica. Universidade Federal de Viçosa. Minas<br />

Gerais. Brasil.<br />

Shlens, J. 2005. A Tutorial on Principal Component Analysis. Center of Neural Science, New<br />

York University. New York City.<br />

Soares, A. 2006. Geoestatística para as ciências da Terra e do Ambiente – 2ª Edição. Instituto<br />

<strong>Superior</strong> Técnico. Lisboa. Portugal.<br />

Zhow, F. Huai-Cheng, G., Yun-Shan, H., Chao-Zhong, W. 2007. Scientometric analysis of<br />

geostatistics using multivariate methods. Scientometrics in press. Budapest <strong>and</strong> Springer,<br />

Dordrecht.<br />

Acknowledgement<br />

Authors would like to express their acknowledgements to Fundação para a Ciência e Tecnologia<br />

(FCT), project REEQ/1163/AGR/2005, CITAB – UTAD, the Portuguese Institute for Nature<br />

Conservation (ICN – Instituto da Conservação da Natureza) <strong>and</strong> the National Association for<br />

Nature Conservation (Quercus - Associação Nacional de Conservação da Natureza).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Etemad & N. Avani 2010. Investigation on species diversity by inventory in Caspian <strong>Forest</strong>s<br />

211<br />

Investigation on species diversity by inventory in Caspian <strong>Forest</strong>s<br />

(Case study: Gorazbon District- Noshahr, Iran)<br />

Vahid Etemad & Nazi Avani<br />

<strong>Forest</strong>ry Group, Faculty of Natural resources, Tehran University, Iran<br />

Abstract<br />

Species diversity is an index for forest ecosystems <strong>and</strong> is an important matter in plant cover<br />

ecology. In this study after %100 inventory in Gorazbon district (Kheyrud <strong>Forest</strong>) <strong>and</strong><br />

establishment of forest type, biodiversity indexes were applied. The results illustrated that the<br />

highest level of the plant species richness, diversity <strong>and</strong> evenness indexes belonged to Carpinus-<br />

Alnus type with Acer sp. <strong>and</strong> the least level of the plant species richness belonged to Fagus <strong>and</strong><br />

Fagus-Carpinus type. The least level of diversity <strong>and</strong> evenness belonged to Fagus <strong>and</strong><br />

Carpinus-Quercus. The results show that sampling is suitable for biodiversity evaluation too.<br />

Keywords: Kheyrud <strong>Forest</strong>, diversity, species richness <strong>and</strong> evenness, Noshahr, Iran<br />

Introduction<br />

Hyrcanus forests with special genetic diversity as number of wooden species <strong>and</strong> plant<br />

communities is very rich <strong>and</strong> have different forest communities (Asadollahi, 2000). The studies<br />

of biodiversity indicators in Hyrcanus forests for investigation on species diversity <strong>and</strong> exact<br />

recognize <strong>and</strong> function of management plan is very necessary.<br />

Moderate regions forests have the main role in biodiversity conservation. Wooden species are<br />

the major matters in the forest ecosystems that have most traces on the other beings life, as<br />

decrease of species diversity can lead weakness other beings.<br />

In biodiversity study, many experts work on it <strong>and</strong> show many formula, such as Sympson<br />

biodiversity indicator in 1949, Manhinic richness indicator in 1964.<br />

The purpose of this study is determination of species diversity in North forests of Iran with<br />

100% inventory as sustainable management of st<strong>and</strong>s under utilization. The results can<br />

investigate status of wooden species as number <strong>and</strong> abundance.<br />

Eshagh Nimvari <strong>and</strong> et al., (2006), for studying of biodiversity in Fagetum community,<br />

Carpino-fagetum <strong>and</strong> Querco-carpinetum in Namkhane <strong>and</strong> Gorazbon district (Noshahr),<br />

richness, species diversity <strong>and</strong> Evenness indicators were measured. The results show that the<br />

most richness, diversity <strong>and</strong> Evenness were in Querco-carpinetum <strong>and</strong> the least was in Fagetum<br />

community. Species diversity in mixed communities was more than pure communities.<br />

Hauk (2005), richness <strong>and</strong> species diversity were studied in Austrian forests <strong>and</strong> the results<br />

showed that richness <strong>and</strong> species diversity in the whole broadleaf communities except Fagetum<br />

was more than conifers. Quercus communities had the most richness , Picea <strong>and</strong> Fagus<br />

communities had the least.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Etemad & N. Avani 2010. Investigation on species diversity by inventory in Caspian <strong>Forest</strong>s<br />

212<br />

Porbabaee (2000), diversity in Fagus habitats in Gilan province is studied <strong>and</strong> shows that in<br />

Fagus communities because of prevailing of Fagus population to another species, species<br />

diversity is very low is this community.<br />

Porbabaee (1999), richness, diversity <strong>and</strong> evenness indicator were studied in forests of Gilan<br />

province, <strong>and</strong> the results show that the least diversity indicator was in Fagus habitats.<br />

Materials <strong>and</strong> methods<br />

Study area<br />

Kheyrud educational <strong>and</strong> investigational forest is located in eastern of Noshahr in North of Iran.<br />

This forest has 8 districts that located in different latitude.<br />

We selected Gorazbon district as 100% inventory was done in this district in the past.<br />

Study method<br />

Sampling<br />

For the first time in autumn of 2007, 100% inventory as execution of selection method of<br />

silviculture in Gorazbon district was suggested <strong>and</strong> done. In this inventory method the whole of<br />

trees (up than 7.5 cm in diameter) were measured. In this method the number of trees <strong>and</strong> shrubs<br />

in each diameter class <strong>and</strong> in each parsel was numbered. With the number of trees in parsels, the<br />

typology map was applied.<br />

Research method<br />

For investigation of biodiversity indicators, we must know about type <strong>and</strong> number of wooden<br />

species that as taken with 100% inventory forms that was done for the entire of forest.<br />

For investigation of biodiversity, there are many indicators but we use of common indicators for<br />

calculation of species diversity. After extraction of data of inventory forms, 8 types in Gorazbon<br />

district were defined. In these types with formulas, Shannon Wiener <strong>and</strong> Sympson diversity<br />

indicators <strong>and</strong> Sympson Evenness indicators was calculated.<br />

For calculation of these species diversity indicators, Ecological methodology software was used.<br />

For drawing of graphs Excel software was used too.<br />

Results<br />

The results show that the most richness is in Carpinus- fagus with Alnus type <strong>and</strong> Carpinus-<br />

Alnus with Acer. The least of richness is in Fagus <strong>and</strong> Fagus- Carpinus type (figure 1).<br />

Shanon <strong>and</strong> sympson species diversity indicators is the most in Carpinus- Alnus with Acer <strong>and</strong><br />

the least in Fagus type (Figure 2, 3).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Etemad & N. Avani 2010. Investigation on species diversity by inventory in Caspian <strong>Forest</strong>s<br />

213<br />

Sympson Evenness indicator is the most in Carpinus - Alnus with Acer type, <strong>and</strong> show that<br />

frequency of species in this type is evenness. This indicator in Fagus type <strong>and</strong> Carpinus –<br />

Quercus is the least (Figure 4).<br />

Conclusion<br />

In recent years, biodiversity <strong>and</strong> climate change are tow subjects that are the major cases in<br />

environment circles. The condition of universe biodiversity is very critical that one of the tow<br />

environments intricate in the universe is biodiversity (Majnonian, 1996).<br />

In a forest, study of st<strong>and</strong> structure can has the key role in biodiversity indicator description<br />

(Franklin, 1993).<br />

As we saw in this study biodiversity indicators: richness, evenness <strong>and</strong> species diversity was the<br />

most in Carpinus - Alnus with Acer type that it was the same to the Eshagh Nimvari (2006)<br />

research.<br />

Statistics results showed that Fagus <strong>and</strong> Fagus- Carpinus types had the least richness indicators<br />

that the same to the results of Hauk (2005) <strong>and</strong> porbabaee (1999 & 2000).<br />

References<br />

Asadollahi, F. 2000, Study species societies in Hircanus region.The Collection of Papers in<br />

National Conference Management of North <strong>Forest</strong>s <strong>and</strong> Sustainable development. Gostare<br />

Population: 323-345.<br />

Ehsagh nimvari, J., Zahedi Amiri, Gh., Marvi Mohajer, M. R., Asadi, M. <strong>and</strong> metaji, A. 2006.<br />

Evaluation <strong>and</strong> comparison of species diversity in fagetum orientalis, Carpino- fagetum<br />

orientalis, <strong>and</strong> Querco carpinetum betulii communities. (Case study: Namkhane <strong>and</strong><br />

Gorazbon Districts- Noshahr). Iranian Journal of <strong>Forest</strong> <strong>and</strong> Poplar Research. Vol. 14, No:<br />

4, 326-337.<br />

Franklin, J. F. 1993. Preserving biodiversity: species ecosystems or l<strong>and</strong>scape Ecological<br />

application. 3: 202-205.<br />

Hauk, E., 2005. Austrian <strong>Forest</strong> Inventory 2000-2002: <strong>Forest</strong> type, Federal Research <strong>and</strong> Training<br />

Center for <strong>Forest</strong>s, Department of <strong>Forest</strong> Inventory, first publication, http://Waldwissen.net.<br />

Majninian, H. 1996. Biodiversity as a key for development. Environment Journal. 8th year, No: 2.<br />

2-7.<br />

Porbabaee, H. 2000. Study of woody species biodiversity in Gilan Fagus <strong>Forest</strong>s. Collection of<br />

Papers National Conference Management of North <strong>Forest</strong>s <strong>and</strong> Sustainable development.<br />

Gostare Population: 51-66.<br />

Porbabaee, H. 1999. Biodiversity of woody plants <strong>and</strong> those Ecosystems in Gilan Provinces. PhD.<br />

These. Tarbiat Modares University.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Etemad & N. Avani 2010. Investigation on species diversity by inventory in Caspian <strong>Forest</strong>s<br />

214<br />

Figure 1: richness indicator in different types<br />

Figure 2: sympson diversity indicator in different types<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Etemad & N. Avani 2010. Investigation on species diversity by inventory in Caspian <strong>Forest</strong>s<br />

215<br />

Figure 3: shanon diversity indicator in different types<br />

Figure 4: sympson evenness indicator in different types<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.E. Eycott et al. 2010. The impact of the matrix on species movement: systematic review <strong>and</strong> meta-analysis<br />

215<br />

The impact of the matrix on species movement: systematic review <strong>and</strong><br />

meta-analysis<br />

A.E. Eycott 1* , G.B. Stewart 2 , G. Br<strong>and</strong>t 1 , L. M. Buyung-Ali 2 , D. E. Bowler 2 , K.<br />

Watts 1 & A.S. Pullin 2<br />

1 <strong>Forest</strong> Research, Alice Holt, Farnham, UK.<br />

2 Centre for Evidence-Based Conservation, School of the Environment <strong>and</strong> Natural<br />

Resources, Bangor University, Gwynedd, UK<br />

Abstract<br />

An increasing number of studies have demonstrated an impact of matrix structure on species<br />

movement. However, the strength of these effects are variable, with no general indications of<br />

which kinds of matrix might be more permeable. In this study we test four possible reasons for<br />

the variability in outcomes from 19 different studies on a range of animal taxa. In particular, we<br />

test whether matrix that is more similar in vertical structure to the 'home' habitat increases<br />

individual movement.<br />

Patch emigration rates were extracted from studies where controlled comparisons or 'choice'<br />

experiments compared the impact of different matrix types on emigration rates.<br />

Matrix types that have a more similar vertical structure to the 'home' habitat were associated<br />

with increased emigration rates. The planning of forests will need to account for ways of<br />

achieving permeable matrix for both forest species moving across cleared areas <strong>and</strong> open<br />

ground species passing through forests.<br />

Keywords. matrix structure, emigration, patch, connectivity<br />

1. Introduction<br />

It is now widely established that the matrix, i.e. the intervening l<strong>and</strong> cover between patches of<br />

habitat, can have a significant impact on species movement (Debinski, 2006; Franklin &<br />

Lindenmayer, 2009; Ricketts, 2001). Ecological network modelling techniques are developing<br />

rapidly to meet the challenge of incorporating matrix impacts (Saura & Pascual-Hortal, 2007).<br />

Such models are currently a key part in the development of some adaptation strategies to help<br />

species move in response to climate change (Hopkins et al., 2007, p.18). What is less clear,<br />

however, is how to account for the matrix in multi-species ecological networks. The data are<br />

limited by taxonomic coverage <strong>and</strong> by spatial <strong>and</strong> temporal scale. To the best of the authors’<br />

knowledge, there are two quantitative analyses of how the matrix affects the abundance of<br />

different taxa (Prugh et al., 2008; Prevedello & Vieira, 2010) but no analyses of movement.<br />

In this paper we apply systematic review <strong>and</strong> meta-analysis to this problem. We examine the<br />

impacts of matrix structure <strong>and</strong> experimental designs on the outcomes of the different studies.<br />

* Corresponding author. Tel.: +44 1420 526200 - Fax: +44 1420 23653<br />

Email address: amy.eycott@forestry.gsi.gov.uk<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.E. Eycott et al. 2010. The impact of the matrix on species movement: systematic review <strong>and</strong> meta-analysis<br />

216<br />

2. Methodology<br />

2.1 Data search<br />

We used a systematic review technique that was originally developed for conservation <strong>and</strong><br />

environmental management from the medical evidence review model (Pullin & Stewart, 2006).<br />

Systematic review strives to minimise error <strong>and</strong> bias through an exhaustive search of peerreviewed<br />

journal publications, grey literature <strong>and</strong> unpublished research findings.<br />

Full details of the systematic review methodology <strong>and</strong> search terms can be found<br />

atwww.environmentalevidence.org/SR43.html. In summary, we aimed to select all articles<br />

(including grey literature <strong>and</strong> web-published data) that presented original, empirical data on<br />

measured emigration rates from habitat patches where two or more matrix types were directly<br />

compared in a controlled experiment or through a single survey. We excluded studies which<br />

measured emigration indirectly, for example by distribution patterns, or where the patches were<br />

too small to support a single individual (e.g. Castellon & Sieving, 2006). We also were unable<br />

to include those papers from which raw emigration rates could not be extracted <strong>and</strong> the authors<br />

of which did not respond to our request for the original data.<br />

2.2 Data synthesis<br />

Matrix types were classified as ‘more favourable’ or ‘less favourable’. This was decided using<br />

the classification by the author of each study. In all but one case, the matrix type that was<br />

assumed to be more favourable was that most structurally similar to the habitat patch, the<br />

exception (Goodwin & Fahrig, 2002) being more ambiguous.<br />

In order to combine the outcomes of different studies, all the data were converted to a single<br />

measure that describes the magnitude of the effect of the study treatment. In this case the<br />

'treatment' is the matrix type. We used risk ratios as the measure of the size of the effect. The<br />

risk ratio is the multiplication of the risk that occurs in a ‘treated’ group relative to a control<br />

group. In this case, the ‘treatment’ was the more similar matrix <strong>and</strong> the ‘control’ the less similar<br />

matrix. If the chance of movement is reduced by a more favourable matrix, the risk ratio will be<br />

less than one; if it increases the chance of individual movement, the risk ratio will be greater<br />

than one.<br />

‘Number needed to treat’ (NNT) can more intuitively illustrate the effect of a 'treatment'. NNT<br />

is defined as the expected number of individuals who need to receive the experimental<br />

‘treatment’ (the permeable matrix) rather than the control (the less permeable matrix) in order<br />

for one additional individual to either incur (or avoid) an event in a given time frame.<br />

The risk ratio was calculated individually for each homogeneous subgroup in a study, e.g. ‘all<br />

the tests performed on species x over a distance of 250 m’. Calculated risk ratios were pooled<br />

across studies to generate an overall weighted average risk ratio within a r<strong>and</strong>om effects model<br />

(DerSimonian & Laird, 1986). The estimate of heterogeneity (variation in risk among studies)<br />

was taken from the Mantel-Haenszel model.<br />

2.3 Covariate analysis<br />

We looked at the impact of five different factors on the study risk ratio: distance, duration,<br />

contrast, choice <strong>and</strong> taxon. Distance <strong>and</strong> duration were investigated using meta-regression.<br />

Contrast, choice <strong>and</strong> taxon were investigated using subgroup analysis, where the meta-analysis<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.E. Eycott et al. 2010. The impact of the matrix on species movement: systematic review <strong>and</strong> meta-analysis<br />

217<br />

is repeated on just one group of data, e.g. all the birds, to see if there is a difference in mean<br />

effect size to the complete analysis.<br />

A ‘contrast’ covariate was generated to define whether the two matrix features being compared<br />

were structurally different. Matrix type was split into four different categories: wooded, grass<br />

or arable, bare or ‘made ground’ (e.g. concrete), <strong>and</strong> water. The ‘contrast’ was zero if both<br />

matrix features were from the same groups <strong>and</strong> one if they came from different groups.<br />

The 'choice' subgroup analysis was determined by the experimental design. Some data were<br />

based on comparisons of emigration from different patches set in different matrix types.<br />

'Choice' was coded zero for these types of experimental design. When the design involved a<br />

comparison of emigration rates through two different matrix types adjacent to different parts of<br />

the edge of the same patch, 'choice' was coded one.<br />

3. Result<br />

We found 19 studies that met our criteria, all of which were animal species across a range of<br />

phyla (birds, fish, insects <strong>and</strong> rodents). Two of these studies were on forest species – both birds<br />

(St. Clair 2003, Desrochers & Hannon 1997). From those 19 studies we extracted 107 data<br />

points that met the homogeneous subgroup criteria. The pooled risk ratio over all studies was<br />

significantly greater than one <strong>and</strong> the lower 95% confidence intervals was not less than one (RR<br />

= 1.32, 95% CI 1.20 to 1.45, p


A.E. Eycott et al. 2010. The impact of the matrix on species movement: systematic review <strong>and</strong> meta-analysis<br />

218<br />

The analysis suggests that permeability is, for many species, related to the structural complexity<br />

of the matrix. It is possible that organisms are adapted to move through structurally similar<br />

habitat types. This could be due to locomotive adaptation or for predator avoidance as species<br />

may avoid different matrix types because they are more conspicuous. This effect was not wholly<br />

consistent across all studies <strong>and</strong> risk ratios below one were found in a number of cases. Some<br />

species may prefer to move through a matrix type structurally dissimilar to their home habitat<br />

for a range of reasons, for example features could be used as visual cues or cover from predators.<br />

While the high heterogeneity <strong>and</strong> correlation of the covariates means further inferences from the<br />

results remain speculative. However, the contrast subgroup analysis may suggest that, in order<br />

to encourage a greater number of individuals to move out of their habitat patch, the matrix may<br />

need to be radically altered to become much more similar to the focal patch habitat structure.<br />

The implication of this would be that a large conservation investment may be necessary to have<br />

a significant effect on connectivity. The 'choice' subgroup analysis may suggest that adding a<br />

permeable route out of a patch is enough to encourage emigration, rather than needing to<br />

surround the patch entirely within a permeable matrix. Routes out of a patch would still need to<br />

be wide enough to provide the matrix required without edge impacts <strong>and</strong> also not lead to lethal<br />

cul-de-sacs (Simberloff et al., 1998).<br />

This review suggests that planning for ecological networks can be generalised to support groups<br />

of species by providing routes through the matrix of similar vegetation structure to their habitat.<br />

Not enough studies were on forest species to compare forest to open-habitat species. In the<br />

majority of densely inhabited regions, altered for hundreds of years by agricultural activity,<br />

fully functional l<strong>and</strong>scapes need to include habitat networks which include permeable elements<br />

for both forest <strong>and</strong> open ground species.<br />

Study<br />

Baum et al., 2004<br />

Bhattacharya et al., 2003<br />

Cronin & Haynes, 2004<br />

DeMaynadier & Hunter, 1999<br />

Desrochers & Hannon, 1997<br />

Goodwin & Fahrig, 2002<br />

Grez, 1997<br />

Grez & Prado, 2000<br />

Haynes & Cronin, 2003<br />

Haynes & Cronin, 2006<br />

Haynes et al., 2007a<br />

Haynes et al., 2007b<br />

Jonsen & Taylor, 2000<br />

Malmgren, 2002<br />

Pither & Taylor, 1998<br />

Ries & Debinski, 2001<br />

Rothermel & Semlitsch, 2002<br />

Russell et al., 2007<br />

Schaefer et al., 2003<br />

St Clair, 2003<br />

Overall (95% CI)<br />

.132872 1 7.52605<br />

Risk ratio<br />

Figure 1: Mean risk ratios from each study. The vertical dashed line represents the mean risk ratio.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.E. Eycott et al. 2010. The impact of the matrix on species movement: systematic review <strong>and</strong> meta-analysis<br />

219<br />

References<br />

Baum, K.A.; Haynes, K.J.; Dillemuth, F.P. & Cronin, J.T. (2004) The matrix enhances the<br />

effectiveness of corridors <strong>and</strong> stepping stones. Ecology, 85: 2671-76.<br />

Bhattacharya, M.; Primack, R.B. & Gerwein, J. (2003) Are roads <strong>and</strong> railroads barriers to<br />

bumblebee movement in a temperate suburban conservation area Biological<br />

Conservation, 109: 37-45.<br />

Castellon, T.D. & Sieving, K.E. (2006) An experimental test of matrix permeability <strong>and</strong><br />

corridor use by an endemic understory bird. Conservation Biology, 20: 135-45.<br />

Cronin, J.T. & Haynes, K.J. (2004) An invasive plant promotes unstable host-parasitoid patch<br />

dynamics. Ecology, 85: 2772-82.<br />

Debinski, D.M. (2006) <strong>Forest</strong> fragmentation <strong>and</strong> matrix effects: the matrix does matter. Journal<br />

of Biogeography, 33: 1791-1792.<br />

deMaynadier, P.G. & Hunter, M.L. (1999) <strong>Forest</strong> canopy closure <strong>and</strong> juvenile emigration by<br />

pool-breeding amphibians in Maine. Journal of Wildlife Management, 63: 441-450.<br />

DerSimonian, R. & Laird, N. (1986) Meta-analysis in clinical trials. Controlled Clinical Trials:<br />

7, 177-188.<br />

Desrochers, A. & Hannon, S.J. (1997) Gap crossing decisions by forest songbirds during the<br />

post-fledging period. Conservation Biology, 11: 1204-1210.<br />

Franklin, J.F. & Lindenmayer, D.B. (2009) Importance of the matrix in maintaining biological<br />

diversity. PNAS, 106: 349-350.<br />

Goodwin, B.J. & Fahrig, L. (2002) How does l<strong>and</strong>scape structure influence l<strong>and</strong>scape<br />

connectivity Oikos, 99: 552-570.<br />

Grez, A.A. (1997) Effect of habitat subdivision on the population dynamics of herbivorous <strong>and</strong><br />

predatory insects in central Chile. Revista Chilena De Historia Natural, 70: 481-490.<br />

Grez, A.A. & Prado, E. (2000) Effect of plant patch shape <strong>and</strong> surrounding vegetation on the<br />

dynamics of predatory coccinellids <strong>and</strong> their prey Brevicoryne brassicae (Hemiptera :<br />

Aphididae). Environmental Entomology, 29: 1244-1250.<br />

Haynes, K.J. & Cronin, J.T. (2003) Matrix composition affects the spatial ecology of a prairie<br />

planthopper. Ecology, 84: 2856-2866.<br />

Haynes, K.J. & Cronin, J.T. (2006) Interpatch movement <strong>and</strong> edge effects: the role of<br />

behavioral responses to the l<strong>and</strong>scape matrix. Oikos, 113: 43-54.<br />

Haynes, K.J.; Diekotter, T. & Crist, T.O. (2007a) Resource complementation <strong>and</strong> the response<br />

of an insect herbivore to habitat area <strong>and</strong> fragmentation. Oecologia, 153: 511-520.<br />

Haynes, K. J., Dillemuth, F. P., Anderson, B. J., Hakes, A. S., Jackson, H. B., Jackson, S. E. <strong>and</strong><br />

Cronin, J. T. (2007b). L<strong>and</strong>scape context outweighs local habitat quality in its effects on<br />

herbivore dispersal <strong>and</strong> distribution. Oecologia 151: 431-441.<br />

Hopkins, J.J., Allison, H.M., Walmsley, C.A., Gaywood, M. & Thurgate, G. (2007) Conserving<br />

biodiversity in a changing climate: guidance on building capacity to adapt. Department<br />

for the Environment, Food <strong>and</strong> Rural Affairs, London.<br />

Jonsen, I.D. & Taylor, P.D. (2000) Fine-scale movement behaviors of calopterygid damselflies<br />

are influenced by l<strong>and</strong>scape structure: an experimental manipulation. Oikos, 88: 553-562.<br />

Malmgren, J.C. (2002) How does a newt find its way from a pond Migration patterns after<br />

breeding <strong>and</strong> metamorphosis in great crested newts (Triturus cristatus) <strong>and</strong> smooth newts<br />

(T.vulgaris). Herpetological Journal, 12: 29-35.<br />

Pither, J. & Taylor, P.D. (1998) An experimental assessment of l<strong>and</strong>scape connectivity. Oikos,<br />

83: 166-174.<br />

Prevedello, J. A. <strong>and</strong> Vieira, M. V. (2010). Does the type of matrix matter A quantitative<br />

review of the evidence. Biodiversity Conservation 19: 1205-1223.<br />

Prugh, L.; Hodges, K.; Sinclair, A. & Brashares, J. (2008) Effect of habitat area <strong>and</strong> isolation on<br />

fragmented animal populations. PNAS, 105: 20770-20775.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.E. Eycott et al. 2010. The impact of the matrix on species movement: systematic review <strong>and</strong> meta-analysis<br />

220<br />

Pullin, A.S. & Stewart, G.B. (2006) Guidelines for systematic review in conservation <strong>and</strong><br />

environmental management. Conservation Biology, 20: 1647-1656.<br />

Ricketts, T.H. (2001) The matrix matters: Effective isolation in fragmented l<strong>and</strong>scapes.<br />

American Naturalist, 158: 87-99.<br />

Ries, L. & Debinski, D.M. (2001) Butterfly responses to habitat edges in the highly fragmented<br />

prairies of Central Iowa. Journal of Animal Ecology, 70: 840-852.<br />

Rothermel, B.B. & Semlitsch, R.D. (2002) An experimental investigation of l<strong>and</strong>scape<br />

resistance of forest versus old-field habitats to emigrating juvenile amphibians.<br />

Conservation Biology, 16: 1324-1332.<br />

Russell, R.E.; Swihart, R.K. & Craig, B.A. (2007) The effects of matrix structure on movement<br />

decisions of meadow voles (Microtus pennsylvanicus). Journal of Mammalogy, 88: 573-<br />

579.<br />

Saura, S. & Pascual-Hortal, L. (2007) A new habitat availability index to integrate connectivity<br />

in l<strong>and</strong>scape conservation planning: Comparison with existing indices <strong>and</strong> application to a<br />

case study. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 83: 91-103.<br />

Schaefer, J.F.; Marsh-Matthews, E.; Spooner, D.E.; Gido, K.B. & Matthews, W.J. (2003)<br />

Effects of barriers <strong>and</strong> thermal refugia on local movement of the threatened leopard<br />

darter, Percina pantherina. Environmental Biology of Fishes, 66: 391-400.<br />

Simberloff, D., Farr, J.A., Cox, J. & Mehlman, D.W. (1992) Movement corridors - conservation<br />

bargains or poor investments. Conservation Biology, 6: 493-504.<br />

St Clair, C.C. (2003) Comparative permeability of roads, rivers, <strong>and</strong> meadows to songbirds in<br />

Banff National Park. Conservation Biology, 17: 1151-1160.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Fernández-Núñez et al. 2010. Eight years of development of a silvopastoral system: effects on floristic diversity<br />

221<br />

Eight years of development of a silvopastoral system: effects on<br />

floristic diversity<br />

E. Fernández-Núñez 1 , A. Rigueiro-Rodríguez 2 & M.R. Mosquera-Losada 2<br />

1 Mountain Research Centre (CIMO), <strong>ESA</strong> – Instituto Politécnico de Bragança, Campus<br />

de Stª Apolónia, Apartado 1172, 53001-855 Bragança, Portugal<br />

2 Crop Production Department, High Politechnic School, University of Santiago de<br />

Compostela, Lugo Campus, 27002-Lugo, Spain<br />

Abstract<br />

Biodiversity is an important issue to promote agricultural sustainability, <strong>and</strong> usually depends on<br />

vegetation management. One of the main reasons to maintain biodiversity is to enhance<br />

productivity in extensive systems, due to the best complementarity between different species to<br />

use soil resources. The objective of this study was to evaluate the effect of two different tree<br />

species, an exotic (Pinus radiata D. Don) <strong>and</strong> a native (Betula alba L.) established at two<br />

densities (833 <strong>and</strong> 2500 tree ha -1 ) <strong>and</strong> three types of fertilization (no fertilization, dairy sewage<br />

sludge fertilization <strong>and</strong> mineral fertilization) on component, species richness <strong>and</strong> abundance<br />

eight years later. The results showed an important reduction in species richness in the systems<br />

established at high density under pine tree compared with birch fertilised with mineral or<br />

without fertilisation. Shannon index was reduced when fertilization was applied under birch at<br />

high density, mostly with dairy sludge, compared with no fertilisation. No effects on plant<br />

diversity was detected when tree density was 833 trees ha -1 .<br />

Keywords: pine, birch, species richness, Shannon index, fertilisation<br />

1. Introduction<br />

In the Northern Spain, where the study was carried out, important changes have occurred in<br />

rural l<strong>and</strong> use in the last decades. The traditional l<strong>and</strong>scape, a heterogeneous mosaic of pastures<br />

<strong>and</strong> native deciduous forests has been gradually substituted by plantations characterized by the<br />

presence of a unique forest tree species (Eucaliptus spp <strong>and</strong> Pinus spp.). The lack of subsequent<br />

management of these plantations has resulted in excessive FCC <strong>and</strong> very high tree densities that<br />

hamper an adequate forest development. Silvopastoral systems may be a viable means to<br />

promote multipurpose forest l<strong>and</strong> use <strong>and</strong> obtain income from newly afforested. Areas managed<br />

for silvopastoralism can reduce fire <strong>and</strong> erosion risk in forests (Rigueiro-Rodríguez et al. 2005),<br />

<strong>and</strong> can enhance biodiversity <strong>and</strong> contribute to the preservation of many endangered species that<br />

depend on ecotones between woodl<strong>and</strong>s <strong>and</strong> open l<strong>and</strong>scapes (Rois-Díaz et al. 2006). However,<br />

these l<strong>and</strong> use changes can cause important modifications to microclimatic conditions (soil,<br />

interior temperature of the system...) <strong>and</strong> this can affect biodiversity in the short <strong>and</strong> medium<br />

term. On the other h<strong>and</strong>, studies carried out in NW Spain have shown that fertilisation enhances<br />

pasture production as well as tree growth in a silvopastoral systems established in very acidic<br />

soils (López-Díaz et al. 2007), but reduces both in neutral soils (Mosquera-Losada et al. 2006)<br />

but, it is important to study how this fertilisation affects to vascular plant biodiversity on a short<br />

time. This study aims to evaluate the effect of Pinus radiata D. Don <strong>and</strong> Betula alba L.<br />

established at two density with different soil fertilisation treatments on alpha biodiversity<br />

over eight years.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Fernández-Núñez et al. 2010. Eight years of development of a silvopastoral system: effects on floristic diversity<br />

222<br />

2. Methodology<br />

2.1 Characteristics of the study site<br />

The experiment was established in Lugo (Galicia, NW Spain) at 439 meters above sea level.<br />

The zone in which the study was carried out is located in the Atlantic Biogeographic Region of<br />

Europe (EEA 2010). The experiment was developed from 1995 to 2003 over a soil classified as<br />

Umbrisol (FAO 1998), with a s<strong>and</strong>y-silty texture (61% s<strong>and</strong>, 34% silt <strong>and</strong> 5% clay) that was<br />

previously used for agricultural purposes (cultivation of potatoes). The initial water pH (1:2.5)<br />

was close to neutrality (6.8), indicating an optimum availability of nutrients for plants (Porta et<br />

al., 2003).<br />

2.2. Experimental design<br />

The experimental design was a completely r<strong>and</strong>omized block with three replicates. In 1995, a<br />

plantation of Pinus radiata D. Don (from container plants) <strong>and</strong> Betula alba L. (from bare root)<br />

were established at 2,500 <strong>and</strong> 833 trees ha -1 . Each experimental unit consisted of a rectangle<br />

square of 5×5 trees with an area of 64 <strong>and</strong> 192 m 2 , respectively. At the end of the winter of<br />

1995, after ploughing, the plots were sown with a mixture of 25 kg ha -1 of Dactylis glomerata L.<br />

var. Saborto, 4 kg ha -1 of Trifolium repens L. var. Ladino <strong>and</strong> 1 kg ha -1 of Trifolium pratense L.<br />

var. Marino. Two types of fertilisation were applied: mineral fertilisation (M) every year<br />

throughout the experiment following a st<strong>and</strong>ard procedure for the region: 500 kg ha -1 of 8:24:16<br />

(N:P 2 O 5 :K 2 O) fertiliser complex in March <strong>and</strong> 40 kg of N (calcium ammonium nitrate 26% N)<br />

ha -1 in May, <strong>and</strong> fertilisation with dairy sludge (D) in the first year (1995) at 154 m 3 ha -1 , i.e.<br />

160 kg of total N, 85.9 kg of total P 2 O 5 <strong>and</strong> 23.4 kg of total K 2 O ha -1 , with values determined<br />

based on total N of sludge. In the two years following (1996 <strong>and</strong> 1997), the plots to which the<br />

sludge was applied were not fertilised, but they were fertilised again from 1998 until the<br />

conclusion of the study (2003), using the same fertilisers <strong>and</strong> doses as for M. One control<br />

treatment was also included in the comparison: no fertilizer (NF). A low pruning (at 2-m high)<br />

was performed on pine at the end of 2001 <strong>and</strong> the birch was given a formational pruning with<br />

the objective of producing quality timber. In this paper, we show the results obtained eight years<br />

after establishment (2003).<br />

2.3 Field samplings<br />

Pasture was harvested in July <strong>and</strong> December under pine at high density <strong>and</strong> on June, July <strong>and</strong><br />

December on the other systems. At each harvest in each experimental unit, the entire surface<br />

area of the nine trees in the centre of the plot delimited by six trees was cleared, the fresh forage<br />

was weighed in situ, <strong>and</strong> a representative subsample was taken to the laboratory. In the<br />

laboratory, the species of 100 g subsamples from each plot were h<strong>and</strong>-separated to determine<br />

botanical composition. The different species were separately weighed to determine dry weight<br />

(48 h at 60 ºC) <strong>and</strong> their relative percentage proportion (taking into account the four harvests of<br />

the year) was calculated. The relative abundance of each species was obtained <strong>and</strong> abundance<br />

diagrams were produced, ordering the species from most to least abundance (Magurran 2004).<br />

Species richness (SR) (Magurran 2004) <strong>and</strong> Shannon–Wiener index (H´) (Shannon <strong>and</strong> Weaver<br />

1949) were also determinated.<br />

2.4 Statistical analyses<br />

The results obtained were analysed with ANOVA following the model: Y ijk = µ + F i +Sp j + B k +<br />

ε ijk , where Y ijk is the studied variable; µ the variable mean; F i : fertilisation; Sp j : tree species; B k :<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Fernández-Núñez et al. 2010. Eight years of development of a silvopastoral system: effects on floristic diversity<br />

223<br />

the block; <strong>and</strong> ε ijk is the error. The LSD test was used for subsequent pairwise comparisons (P <<br />

0.05; α= 0.05). The statistical software package SAS (2001) was used for all analyses.<br />

3. Results<br />

26 species were found eight years after establishment, belonging to 9 different families. Eight<br />

species belonged to the family Asteraceae (31%), 7 to the Poaceae (27%), 3 to the<br />

Leguminosae, 2 to the Geraniaceae, 2 to the Polygonaceae, leaving just one representative for<br />

each of the families Caryophyllaceae, Umbelliferae, Plantaginaceae <strong>and</strong> Rosaceae (see Figure<br />

1). Of these, 54% are perennial species, 42% are annual species <strong>and</strong> 4% are biannual. The total<br />

number of species (SR) under pine was 2, 5 <strong>and</strong> 1 at 2,500 trees ha -1 <strong>and</strong> 6, 12 <strong>and</strong> 13 at 833<br />

trees ha -1 for the treatments D, M <strong>and</strong> NF, respectively. Under birch, SR was 5, 10 <strong>and</strong> 17 <strong>and</strong> 8,<br />

8 <strong>and</strong> 14 at high <strong>and</strong> low density <strong>and</strong> D, M <strong>and</strong> NF treatments, respectively. Dactylis glomerata<br />

was the most abundant species in all treatments evaluated; Dactylis species was above 70%<br />

under pine <strong>and</strong> 50% under birch, independently of density. Holcus lanatus L. <strong>and</strong> Agrostis<br />

capillaris L. appeared also in all treatments, with the exception of those systems established<br />

with pine at high density <strong>and</strong> fertiliser with dairy sludge (D) <strong>and</strong> no fertiliser (NF). On the<br />

contrary, there were species, all annuals, that were only associated to certain treatments e.g<br />

Erodium moschatum (L.) L´Hér <strong>and</strong> Lolium multiflorum Lam under pine at low density or<br />

Bromus di<strong>and</strong>rus Roth, Chamaemelum mixtum L., Conyza canadensis L., Cerastium<br />

glomeratum Thuill <strong>and</strong> Sonchus oleraceus L. that were cited only under birch systems (the first<br />

three at high density <strong>and</strong> the last at low density). Finally, Shannon index (H´) was reduced when<br />

fertilization was applied under birch at high density, mostly with dairy sludge (D), compared<br />

with control (NF) No effects of tree species or fertilization on plant diversity was detected when<br />

tree density was 833 trees ha -1 .<br />

4. Discussion<br />

It is generally assumed that even-aged forest plantations are negative from the viewpoint of<br />

biodiversity conservation because, they are characterized by loss of horizontal <strong>and</strong> vertical<br />

heterogeneity, fundamental components of forest biodiversity (Marcos et al 2007). Our results<br />

show that this is true only, if at the moment of establishment of the system, the choice of tree<br />

cover <strong>and</strong> tree density are not adequate. In the year of establishment (1995), when the<br />

development of canopy was small <strong>and</strong> the understory environmental conditions were quite<br />

similar between both tree systems, SR was very similar between them (23 <strong>and</strong> 20 species under<br />

high <strong>and</strong> lower density, respectively, of which approximately 55% were annuals) (Fern<strong>and</strong>ez-<br />

Núñez et al 2010). From the beginning of study (1995), the growth of pine was greater than that<br />

of birch (Rigueiro-Rodríguez et al. 2000) <strong>and</strong> resulted in a faster canopy closure in pine than<br />

birch in the systems established at high density five years after establishment (year 2000)<br />

(Fernández-Núñez et. al 2010). This canopy closure caused a deep shade, especially in spring,<br />

lead to fast reductions of vascular plants when biodiversity is compared among the years 1995<br />

<strong>and</strong> 2003 (88% vs 49% at high density under pine <strong>and</strong> birch, respectively <strong>and</strong> 48% vs 52% at<br />

low density, under pine <strong>and</strong> birch, respectively). This reduction effect was especially important<br />

on annuals species, as they dissapiared under pine at high density <strong>and</strong> reduced their proportion<br />

around 67% in the other systems, which depend on high levels of radiation, that in our case, was<br />

mostly intercepted by pine at a high density. Moreover, the higher accumulation of litter <strong>and</strong><br />

humus under pine at high density (Mosquera-Losada et al. 2006b), that is know to contain plant<br />

growth inhibitors (Piggot 1990), contributed to enhance biodiversity losses. On the other h<strong>and</strong>,<br />

a negative relationship between mineral fertilizer (M) <strong>and</strong> development of pine at hight<br />

density (Mosquera-Losada et al. 2006) was found from the begining of study. This<br />

effect has been repeated until year 2003 (data not presented), <strong>and</strong> results in a lower<br />

reduction on alpha biodiversity. While, under birch at hight density, alpha biodiversity<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Fernández-Núñez et al. 2010. Eight years of development of a silvopastoral system: effects on floristic diversity<br />

224<br />

was reduced when fertiliser was applicated (D or M). These treatments have favoured<br />

the presence of Dactylis glomerata (80, 60, 39% under D, M <strong>and</strong> NF, respectively).<br />

This gramineae is characterised by strong growth in height that limit the presence of<br />

other species (Rodríguez et al. 2001). At low density, SR was reduced by D treatment<br />

around 50% when compared with M y NF treatments in pine st<strong>and</strong>s. Dactylis glomerata<br />

<strong>and</strong> Rumex obtusifolius L. were the most abundant species in D. Some Polygonaceae<br />

plants including Rumex spp. have been described as allelopathic <strong>and</strong> reduced the<br />

germination of Lolium spp (Lutts et al. 1987) <strong>and</strong> inhibited some grasses <strong>and</strong> controlled<br />

their spatial distribution (Carballeria et al. 1988). Regarding birch established at low<br />

density, D <strong>and</strong> M treatments had lower biodiversity than NF in spite of being birch<br />

smaller in those treatments compared with NF. Both D <strong>and</strong> M had an important<br />

proportion of Dactylis glomerata, Agrostis capillaris L. <strong>and</strong> Holcus lanatus L. Tree<br />

canopy develop also caused a reduction on evenness <strong>and</strong> therefore an increase of the dominance<br />

of perennial species (Dactylis glomerata, Holcus lanatus <strong>and</strong> Agrostis capillaris), irrespective<br />

of the overstorey tree species <strong>and</strong> density.<br />

References<br />

Carballeira, A., E. Carral <strong>and</strong> M. Reigosa. 1988. Asymmetric scale distribution <strong>and</strong> allelopathy:<br />

interaction between Rumex obtusifolius <strong>and</strong> meadow species. Jounal of Chemical<br />

Ecology, 14: 1775-1785.<br />

EEA 2010 http://www.eea.europa.eu/data-<strong>and</strong>-maps/figures/biogeographical-regions-in-europe<br />

Ellsworth, J., Harrington, R. <strong>and</strong> Fownes, J., 2004. Seedling emergence, growth, <strong>and</strong> allocation<br />

of oriental bittersweet: effects of seed input, seed bank, <strong>and</strong> forest floor litter. <strong>Forest</strong><br />

Ecology <strong>and</strong> Management 190: 255-264.<br />

FAO-ISRIC-ISSS, 1998. World Referent Base for Soil Resources. World Soil Resources<br />

Reports 84, Rome.<br />

Fernández-Núñez, E., Rigueiro-Rodríguez, A. <strong>and</strong> Mosquera-Losada, M.R., 2010. Afforestation<br />

of agricultural l<strong>and</strong> with Pinus radiata D. Don <strong>and</strong> Betula alba L. in NW Spain: Effects<br />

on soil pH, understorey production <strong>and</strong> floristic biodiversity eleven years after<br />

establishment. L<strong>and</strong> Degradation <strong>and</strong> Development (in press).<br />

Porta Casanellas, J., Roquero De Laburo, C. <strong>and</strong> López-Acevedo, M., 2003. Edafología para la<br />

agricultura y el medio ambiente. Mundi Prensa, Madrid, 959 p.<br />

Rigueiro-Rodríguez, A., Mosquera-Losada, M.R., Romero- Franco, R., González-Fernández,<br />

M.P. <strong>and</strong> Villarino-Urtiaga, J.J., 2005. Silvopastoral systems as forest fire prevention<br />

technique. In: MR Mosquera-Losada, J McAdam <strong>and</strong> A Rigueiro-Rodríguez (Eds.)<br />

Silvopastoralism <strong>and</strong> sustainable l<strong>and</strong> management, UK: CAB Internacional: 380-387.<br />

López-Díaz, M.L., Mosquera-Losada, M.R. <strong>and</strong> Rigueiro-Rodríguez, A., 2007. Lime, sewage<br />

sludge <strong>and</strong> mineral fertilization in a silvopastoral system developed in very acid soils.<br />

Agroforestry Systems, 70 (1): 91–101.<br />

Lutts, S., Peeters, A. <strong>and</strong> Lambert, J., 1987. Contribution to the study of allelopathy in Rumex<br />

obtusifolius L. Bull. Soc. Royle de Botanique de Belgique, 120: 143-152.<br />

Magurran, A.E., 2004. Measuring biological diversity. Blackwell Publishing, Malden,<br />

Massachusetts, 256 p.<br />

Mosquera-Losada, M.R., Fernández-Núñez, E. <strong>and</strong> Rigueiro-Rodríguez, A., 2006. Pasture, Tree<br />

<strong>and</strong> Soil Evolution in Silvopastoral Systems of Atlantic Europe. <strong>Forest</strong> Ecology <strong>and</strong><br />

Management, 232: 135-145.<br />

Mosquera-Losada, M.R., Fernández-Núñez, E. <strong>and</strong> Rigueiro-Rodríguez, A., 2006b.<br />

Contribución de las acículas al pasto tras la poda baja y aporte en nutrientes en un<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Fernández-Núñez et al. 2010. Eight years of development of a silvopastoral system: effects on floristic diversity<br />

225<br />

sistema silvopastoral desarrollado con Pinus radiata. Actas del IV Congreso Nacional<br />

<strong>Forest</strong>al. Zaragoza.<br />

Marcos, J.A., Marcos, E., Taboada, A. <strong>and</strong> Tárrega, R., 2007. Comparison of community<br />

structure <strong>and</strong> soil characteristics in different aged Pinus sylvestris plantations <strong>and</strong> a<br />

natural pine <strong>Forest</strong>. <strong>Forest</strong> Ecology <strong>and</strong> Management, 247(1-3):35-42.<br />

Piggot, C.D., 1990. The influence of evergreen coniferous nurse-crops on the field layer on two<br />

woodl<strong>and</strong> communities. Journal of Applied Ecology 27, 448-459.<br />

Rigueiro-Rodríguez, A., Mosquera-Losada, M.R. <strong>and</strong> Gatica-Trabanini, E., 2000. Pasture<br />

production <strong>and</strong> tree growth in a young pine plantation fertilized with inorganic fertilizers<br />

<strong>and</strong> milk sewage in northwestern Spain. Agroforestry Systems, 48: 245–256.<br />

Rodríguez, M., Gómez-Sal, A., García, R., Moro, A. <strong>and</strong> Calleja, A., 2001. Relaciones entre la<br />

producción, diversidad y riqueza de especies en prados fertilizados. Pastos: 175-180.<br />

Rois-Díaz, M., Mosquera-Losada, R. <strong>and</strong> Rigueiro-Rodriguez, A., 2006. Biodiversity Indicators<br />

on Silvopastoralism across Europe. European <strong>Forest</strong> Institute, Joensuu, Finl<strong>and</strong>.<br />

SAS 2001. SAS/Stat User´s Guide: Statistics. SAS Institute Inc. Cary, NC, USA.<br />

Shannon, C.E. <strong>and</strong> Weaver, W., 1949. The mathematical theory of communication. Univ.<br />

Illinois Press, Illinois, 116 p.<br />

1.00<br />

0.50<br />

0.00<br />

*Dg(2)<br />

*Hl(2)<br />

*Dg(2)<br />

*Hl(2)<br />

*Ag(2)<br />

*Se(1)<br />

*Hm(2)<br />

*Dg(2)<br />

*Dg(2)<br />

*Hl(2)<br />

*Lp(2)<br />

*Ag(2)<br />

*Ta(1)<br />

*Dg(2)<br />

*Hl(2)<br />

*Ag(2)<br />

**Gd(4)<br />

D(7)<br />

*Ro(5)<br />

**Br(2)<br />

**Co(1)<br />

**Ce(6)<br />

*Rub(1)<br />

*Dg(2)<br />

*Ag(2)<br />

**Tc(3)<br />

*Hl(2)<br />

*Rub(1)<br />

**Br(2)<br />

D(7)<br />

*Ac(1)<br />

*Hm(2)<br />

*Tr(3)<br />

**Ch(1)<br />

**Ce(6)<br />

*Lo(3)<br />

**Mat(1)<br />

*Se(1)<br />

*Lp(2)<br />

*Ta(1)<br />

D M NF D M NF<br />

SR:2 SR:5 SR:1 SR:5 SR:10 SR:17<br />

Pine<br />

Birch<br />

2500 trees ha-1 2500 trees ha-1<br />

1.00<br />

0.50<br />

0.00<br />

*Dg(2)<br />

*Ro(5)<br />

*Ag(2)<br />

D(7)<br />

*Hl(2)<br />

**Lm(2)<br />

*Dg(2)<br />

*Hl(2)<br />

*Ag(2)<br />

*Ta(1)<br />

*Se(1)<br />

**Er(4)<br />

*Tr(3)<br />

D(7)<br />

*Ac(1)<br />

**Cre(1)<br />

**Tc(3)<br />

*Pl(8)<br />

*Dg(2)<br />

*Ag(2)<br />

*Hl(2)<br />

**Tc(3)<br />

*Ta(1)<br />

**Mat(1)<br />

*Se(1)<br />

D(7)<br />

*Pl(8)<br />

**Cre(1)<br />

**Gd(4)<br />

*Rub(9)<br />

*Rac(5)<br />

*Dg(2)<br />

*Ag(2)<br />

*Hl(2)<br />

D(7)<br />

*Ta(1)<br />

*Pl(8)<br />

*Rub(9)<br />

*Rac(5)<br />

*Dg(2)<br />

*Hl(2)<br />

**Cre(1)<br />

D(7)<br />

*Se(1)<br />

*Pl(8)<br />

*Tr(3)<br />

*Ro(5)<br />

*Dg(2)<br />

*Ta(1)<br />

D(7)<br />

**Tc(3)<br />

*Hl(2)<br />

*Se(1)<br />

**Cre(1)<br />

**So(1)<br />

*Ag(2)<br />

*Tr(3)<br />

*Pl(8)<br />

**Ce(6)<br />

**Rac(2)<br />

**Rub(9)<br />

D M NF D M NF<br />

SR: 6 SR: 12 SR: 13 SR:8 SR:8 SR:14<br />

Pine<br />

Birch<br />

833 trees ha-1 833 trees ha-1<br />

Figure 1: Abundance diagrams in the plots established under Pinus radiata D. Don (Pine) <strong>and</strong> Betula alba<br />

L. (Birch) at two densities (2500 <strong>and</strong> 833 trees ha -1 ). Note: SR: species richness, D: fertilised with sludge,<br />

M: mineral fertiliser, NF: not fertilised, Ac: Achillea millefolium L.; Ag: Agrostis capillaris L.; Br:<br />

Bromus di<strong>and</strong>rus Roth; Ce: Cerastium glomeratum Thuill; Ch: Chamaemelum mixtum L.; Co: Conyza<br />

canadensis L.; Cre: Crepis capillaris (L.) Wallr; D: Daucus carota L.; Dg: Dactylis glomerata L.; Er:<br />

Erodium moschatum (L.) L´Hér; Gd: Geranium dissectum L.; Hl: Holcus lanatus L.; Hm: Holcus mollis<br />

L.; Lo: Lotus corniculatus L.; Lm: Lolium multiflorum Lam; Lp: Lolium perenne L.; Mat: Chamomilla<br />

recutita L.; Pl: Plantago lanceolata L.; Rac: Rumex acetosella L.; Ro: Rumex obtusifolius L.; Rub:<br />

Rubus sp.; Se: Senecio jacobaea L.; So: Sonchus oleraceus L.; Ta: Taraxacum officinale Weber; Tc:<br />

Trifolium campestre Schreber <strong>and</strong> Tr: Trifolium repens L.; families: (1) Asteraceae, (2) Poaceae, (3)<br />

Leguminosae, (4) Geraniaceae, (5) Polygonaceae, (6) Caryophyllaceae, (7) Umbelliferae, (8)<br />

Plantaginaceae <strong>and</strong> (9) Rosaceae, (*): perennial species, (**): annual species <strong>and</strong> ( ): biannual species.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Fernández-Núñez et al. 2010. Eight years of development of a silvopastoral system: effects on floristic diversity<br />

226<br />

2<br />

Shannon´s Index (H´)<br />

a<br />

1<br />

bc<br />

ab<br />

0<br />

cd<br />

d<br />

d<br />

D M NF D M NF D M NF D M NF<br />

Pine Birch Pine Birch<br />

2x2<br />

Treatments<br />

3x4<br />

Figure 2: Shannon-Wiener index (H´) determined for each of the fertiliser treatments applied under the<br />

two types of trees <strong>and</strong> densities. Note: D: fertilised with sludge, M: mineral fertiliser <strong>and</strong> NF: not<br />

fertilised. Different letters indicate significant differences between treatments in the same density.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Fracassi & D. Somma 2010. Participatory action research concerning the l<strong>and</strong>scape use by a native cervid in a wetl<strong>and</strong><br />

227<br />

Participatory action research concerning the l<strong>and</strong>scape use by a native<br />

cervid in a wetl<strong>and</strong> of the Plata Basin, Argentina<br />

Natalia Fracassi 1* & Daniel Somma 1,2<br />

1<br />

Delta Research Station-National Institute of Agricultural Technology (INTA),<br />

Argentina<br />

2<br />

National Parks Administration, Argentina<br />

Abstract<br />

The marsh deer is one of the few deer species restricted to wetl<strong>and</strong> habitats. It is considered<br />

“threatened” at both national <strong>and</strong> international levels. In the Delta of the Paraná River part of<br />

the natural vegetation has been converted to industrial forest after to drain the l<strong>and</strong>. Thus, the<br />

persistence of the species at the l<strong>and</strong>scape level may depend on its adaptation <strong>and</strong> the attitude of<br />

the local people towards the deer. From 26 stakeholders’ interviews, we evaluate how the cervid<br />

uses the l<strong>and</strong>scape <strong>and</strong> what threats this species is facing. The deers were registered mostly in<br />

adult <strong>and</strong> young poplar afforestations followed by adult willow afforestations, but interviewees<br />

agreed that afforestations that maintain understory foliage function as refuges for deers. In turn,<br />

interviewees agreed that the deer population declined in the last 10 years <strong>and</strong> current level of<br />

hunting is the biggest problem for the regional marsh deer population.<br />

Keywords: marsh deer, wetl<strong>and</strong>, afforestations, stakeholders, l<strong>and</strong>scape, threats<br />

1. Introduction<br />

The marsh deer (Blastocerus dichotomus, Illiger, 1811) is the largest <strong>and</strong> probably the most<br />

endangered cervid in South America (Thornback <strong>and</strong> Jenkins 1982; Fonseca et al. 1994). The<br />

species is distributed from mid-western <strong>and</strong> southern Brazil, Paraguay, eastern Bolivia <strong>and</strong> a<br />

small portion of south-eastern Peru to northern Argentina (Pinder <strong>and</strong> Grosse 1991; Tomás et al.<br />

1997; Wemmer 1998; D´Alessio et al. 2001). It is considered “threatened” at both national<br />

(Díaz <strong>and</strong> Ojeda 2000) <strong>and</strong> international (IUCN, 2002) levels, <strong>and</strong> it is one of the few deer<br />

species known to be restricted to wetl<strong>and</strong>s such as seasonal streams, swamps, <strong>and</strong> flooded<br />

savannas (Schaller <strong>and</strong> Hamer 1978). In Argentina, the Delta of the Paraná River (provinces of<br />

Buenos Aires <strong>and</strong> Entre Rios) constitutes the austral distribution boundary of the species <strong>and</strong> it<br />

holds one of the three main populations of this cervid in the country (Chébez 1994, D’Alessio et<br />

al. 2001).<br />

The Lower Delta of the Paraná River is a wide costal freshwater wetl<strong>and</strong> characterized by a<br />

vegetation mosaic crossed by an intricate network of rivers. Native forests develop on levees<br />

<strong>and</strong> relatively elevated sites, whereas lowl<strong>and</strong>s are colonised by marshes or rush communities<br />

(K<strong>and</strong>us et al. 2003). A part of the terminal portion of this wetl<strong>and</strong> has been altered in the last<br />

80 years mainly due to the replacement of natural vegetation by forestry plantations after the<br />

drainage of the l<strong>and</strong>. Three different population nuclei of marsh deer have been proposed to<br />

occur in this wetl<strong>and</strong>, one in an undisturbed area covered with native vegetation <strong>and</strong> the two<br />

* Corresponding author. Tel.: 0054-3489-460075 - Fax: 0054-3489-460076<br />

Email address: nfracassi@correo.inta.gov.ar/ natfracassi@yahoo.com.ar<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Fracassi & D. Somma 2010. Participatory action research concerning the l<strong>and</strong>scape use by a native cervid in a wetl<strong>and</strong><br />

228<br />

remaining in areas mostly dominated by forestry plantations (D´Alessio et al. 2002). The<br />

recently created Delta of the Paraná River Biosphere Reserve (MAB-UNESCO) currently<br />

protects the nuclei inhabiting the undisturbed area. The two remaining nuclei, on the other h<strong>and</strong>,<br />

are established in non-protected human-dominated areas where productive activities with low<br />

sustainable management appear to be threatening the long-term persistence of the species.<br />

The population status of the marsh deer in this wetl<strong>and</strong> under current conditions is unknown<br />

(Dellafiore <strong>and</strong> Maceira 1998). So, an evaluation of the main threats that affect this population<br />

<strong>and</strong> the detection of how individuals use the various components of the altered l<strong>and</strong>scape are<br />

necessary. In turn, conservation strategy generated for the marsh deer can also bring<br />

conservation benefits to other wild species <strong>and</strong> the whole Delta as well.<br />

As the persistence of the species at the l<strong>and</strong>scape level may depend on its adaptation to the<br />

altered habitats <strong>and</strong> the attitude of the local people towards the deer, the aim of this study was to<br />

evaluate the local people perception about how this cervid uses the l<strong>and</strong>scape <strong>and</strong> which threats<br />

are affecting the species.<br />

2. Methodology<br />

The study area covers around 500km 2 <strong>and</strong> is centred in the downstream area of the Lower Delta<br />

of the Paraná River region (Buenos Aires province, Argentina). It includes the accretion portion<br />

of the delta between Paraná de las Palmas <strong>and</strong> Paraná Guazú rivers. The Lower Delta region<br />

(2,700km 2 ) has a temperate climate with a mean annual temperature of 289.96K, <strong>and</strong> an annual<br />

rainfall of 1,073 mm. The hydrological regime is the result of the combined effects of the<br />

Paraná river flow <strong>and</strong> the tidal pattern of the Del Plata estuary (Mujica 1979; Minotti et al.<br />

1988). About 50% of the region is affected by human activities, mainly by the development of<br />

Salicaceae plantations (Salix spp. –willow- <strong>and</strong> Populus spp. –poplar-) <strong>and</strong> tourist <strong>and</strong><br />

recreation activities (K<strong>and</strong>us 1997).<br />

This Delta holds the largest Salicaceae plantations in the country (Petray 2000) <strong>and</strong> this activity<br />

constitutes the economically most significant in the region (SAGPyA 1999; Casaubón et al.<br />

2001) with 75% of willow <strong>and</strong> 25% of poplar afforestations (Borodowski 2006). Before<br />

planting, systematic work is carried out through the construction of canals <strong>and</strong> drainage ditches<br />

of different sizes <strong>and</strong> / or construction of embankments along the coast (locally known as<br />

“diques” <strong>and</strong> “atajarrepuntes”), which prevent entry of water from flooding (Casaubón et al.<br />

2001). In the study area the poplar plantations are made for sawing wood, with large planting<br />

spacing ranging from 6x2 m to 6x6 m (Poplar Commission 1985). The willow, on the other<br />

h<strong>and</strong>, is used mainly for paper <strong>and</strong> particle board industries <strong>and</strong> plantations vary in spacing on<br />

3x3 <strong>and</strong> 3x2 m. In the case of poplar, the understory foliage management implies keeping the<br />

soil free of weeds.<br />

The methodology was based on the interview to stakeholders (local people, producers, <strong>and</strong> l<strong>and</strong><br />

managers). Six questions were included in the interview: 1) Area occupied by each habitat type<br />

in the area of influence (e.g. ranch) of the interviewer; 2) Which habitat types are used by deers<br />

to transit <strong>and</strong> feed; 3) Which habitats appear to be used by deers as refuges; 4) Did the<br />

population decline in the last 10 years; 5) Which are the main threats for the deer population;<br />

<strong>and</strong> 6) How can we contribute to the protection of the marsh deer.<br />

Habitat types were divided into: 1) adult willow afforestations; 2) young willow afforestations;<br />

3) adult poplar afforestations; 4) young poplar afforestations; 5) marshes; 6) Cattle pastures; 7)<br />

afforestations without management or ab<strong>and</strong>oned; 8) riparian native forests; 9) roads; <strong>and</strong> 10)<br />

others. A habitat was considered as a refuge for deers when it was used by individuals to rest or<br />

to escape from humans.<br />

3. Results<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Fracassi & D. Somma 2010. Participatory action research concerning the l<strong>and</strong>scape use by a native cervid in a wetl<strong>and</strong><br />

229<br />

During 2009, 26 stakeholders (with an area of influence of 250.6 km2) were interviewed. The<br />

habitat types included in this area of influence are shown in Figure 1. The most representative<br />

habitats were the young <strong>and</strong> adult willow (49.7%) <strong>and</strong> poplar (28.3%) afforestations, whereas<br />

the rest of the habitats occupied a smaller area.<br />

Most of the records were reported in almost all habitats but mainly in adult poplar afforestations<br />

(24.7%) (see Figure 2). On the other h<strong>and</strong>, interviewees reported that ab<strong>and</strong>oned afforestations<br />

or afforestations without understory management were the habitat most used as refuge by deers,<br />

followed by poplar <strong>and</strong> willow afforestations (see Figure 3). In turn, interviewees reported that a<br />

half of the poplar afforestations used as refuge by deers corresponds to afforestations without<br />

understory management or with presence of herbaceous <strong>and</strong> shrubs.<br />

Regarding the population status, almost 60% of interviewees reported that there are fewer deers<br />

in the population than 10 years ago, while the remaining responded that the number of deer is<br />

equal or higher (19% <strong>and</strong> 23%, respectively) than 10 years ago (see Figure 4). Hunting was<br />

reported as the principal threat that affects the species in the area, whereas habitat modification<br />

(ej. embankments construction <strong>and</strong> drainage of l<strong>and</strong>s) <strong>and</strong> the interaction with cattle (ej.<br />

competition <strong>and</strong> transmission of diseases) were considered secondary threats (see Table 1).<br />

Interviewees suggested that the effective control of hunting (by their own or by control agents)<br />

could be the most appropriate action contributing to the conservation of the deer (see Table 2).<br />

In turn, they also considered as an important conservation measure to leave unmanaged<br />

afforestations within their own fields as a refuge for deer. Finally, they agreed that education in<br />

schools <strong>and</strong> awareness within the general public <strong>and</strong> decision makers would help minimize the<br />

risk level of the deer population in the Lower Delta.<br />

4. Discussion<br />

According to interviewees, despite the low representation of natural habitats in the study area<br />

(marshes <strong>and</strong> riparian forests) marsh deer are present <strong>and</strong> uses the new habitats generated by<br />

forest industries as transit <strong>and</strong> feeding sites. However, the greater visibility inside poplar<br />

plantations due to the planting distance <strong>and</strong> the lack of understory foliage could improve the<br />

detection of deers in those areas in comparison with other more closed habitats. Considering the<br />

use of different habitats as refuge, the interviewed agreed that the marsh deer uses the denser<br />

poplar afforestations with presence of understory proportionally more often than the rest of the<br />

poplar because this habitat provides greater protection <strong>and</strong> resting sites. On the other h<strong>and</strong>, both<br />

cattle pastures <strong>and</strong> roads would be less optimal habitat due to the lower presence of refuges.<br />

In spite of the possibility that the species may have adapted to the habitat transformation, the<br />

current level of hunting continues threatening the deer population in the region. The possibility<br />

of diseases transmission from cattle to deers <strong>and</strong> the increase of suboptimal habitats due to the<br />

mish<strong>and</strong>ling of understory foliage <strong>and</strong> water in the afforestations are also threats for this<br />

population. In relation to water management, people perception agreed with Varela (2003), who<br />

determined that the l<strong>and</strong> use type associated with hydrological management is a key factor in<br />

the distribution patterns of marsh deer in afforestations. Periodic openings of the dams<br />

floodgates (or diques) would allow the coexistence of forest plantations with high-rich patches<br />

of forage resources for the deer. Thus, an analysis of the effects of water management inside<br />

afforestations <strong>and</strong> hunting on the population is urgently needed. In addition, the analysis of the<br />

distribution, size, <strong>and</strong> connectivity of optimal <strong>and</strong> suboptimal habitat patches at different<br />

l<strong>and</strong>scape levels could allow an assessment of the regional population viability.<br />

At the afforestations level, is essential to delineate a scheme of plantation planning <strong>and</strong> forest<br />

management that considers relevant aspects such as (1) understory management (at least in<br />

poplar plantations), (2) management of patches in ab<strong>and</strong>oned afforestations, (3) the inclusion of<br />

high density willow plantations in some areas as an alternative to generate favorable refuge<br />

patches for the deer, <strong>and</strong> (4) l<strong>and</strong>scape management of forestry establishments. Therefore, the<br />

detection <strong>and</strong> design of corridors connecting feeding zones with refuge areas in the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Fracassi & D. Somma 2010. Participatory action research concerning the l<strong>and</strong>scape use by a native cervid in a wetl<strong>and</strong><br />

230<br />

afforestations <strong>and</strong> others productive l<strong>and</strong>s would contribute to the persistence of the deer<br />

regional population.<br />

It would be also relevant to evaluate the effects of the cattle production system on the deer<br />

population as well, because the regional increase in cattle numbers related to the expansion of<br />

silvopastoral systems (Arano, 2006) may constitute a new potential threat. This analysis should<br />

include both sanitary <strong>and</strong> competence aspects. Finally, it is essential to organize <strong>and</strong> perform<br />

environmental awareness campaigns focusing on stakeholders <strong>and</strong> decision makers. Based on<br />

the appropriate application of laws <strong>and</strong> the increased knowledge of the ecology <strong>and</strong> demography<br />

of marsh deers, is possible to implement more effective conservation programs. In this context,<br />

the participatory action research approach considered in this study can provide a suitable<br />

platform to build these conservation programs, generating the commitment of the local people<br />

since the beginning of the program execution.<br />

References<br />

Arano, A. 2006. Ganadería en sistemas silvopastoriles del Delta del Paraná. EEA INTA Delta<br />

del Paraná. Monografía. Sitio argentino de Producción Animal.<br />

Borodowski, E., 2006. Álamos y sauces en el Delta del Paraná: situación del sector y<br />

Silvicultura. Actas Jornadas de Salicáceas 2006, 61-70pp.<br />

Chébez, J.C., 1994. Los que se van. Ed. Albatros. Buenos Aires.<br />

Casaubon, E., Gurini, L.B., <strong>and</strong> Cueto, G., 2001. Diferente calidad de estación en una<br />

plantación de Populus deltoides cv. Catfish 2 del bajo Delta Bonaerense del Río Paraná<br />

(Argentina). Invest. Agr.: Sist. Recur. For. Vol 10(2).<br />

D’Alessio, S., Varela, D., Gagliardi, F., Lartigau, B., Aprile, G., Mónaco, G., <strong>and</strong> Heinonen, S.,<br />

2001. El Ciervo de los Pantanos (Blastocerus dichotomus). En: Los Ciervos Autóctonos<br />

de la Argentina y la acción del hombre. (Eds.: C. Dellafiore y N. Maceira). Secretaría<br />

de Desarrollo Sustentable y Política Ambiental. Buenos Aires, 95pp.<br />

D’Alessio, S., Varela, D., Lartigau, B., Gagliardi, F., Aprile, G. <strong>and</strong> Mónaco C., 2002. Marsh<br />

Deer Project. Marsh Deer (Blastocerus dichotomus): A Flagship Species for the Delta of<br />

Paraná <strong>and</strong> it’s Biodiversity. Final Report to BP Conservation Programme.<br />

Dellafiore, C.M. <strong>and</strong> Maceira, N.O., 1998. Problemas de conservación de los ciervos autóctonos<br />

de la Argentina. Mastozoología Neotropical 5(2): 137-145.<br />

Díaz, G.B. <strong>and</strong> Ojeda R.A. (eds.). 2000. Libro rojo de mamíferos amenazados de la Argentina.<br />

Sarem.<br />

Fonseca, G.A., Bda, A.B., Ryl<strong>and</strong>s, C.M.R., Costa, R.B., Machado, <strong>and</strong> Leite, Y.LR., 1994.<br />

Livro Vermelho dos Mamíferos Brasileiros Ameaçados de Extinção. Fundação<br />

Biodiversitas. Belo Horizonte.<br />

IUCN. 2002. 2002 IUCN Red List of Threatened Species. IUCN. Gl<strong>and</strong> y Cambridge.<br />

(Disponible en Intenet www.redlist.org).<br />

K<strong>and</strong>us, P; Malvarez, A. I, <strong>and</strong> Madanes, N.,. 2003. Estudio de las comunidades de plantas<br />

herbáceas de las islas bonaerenses del Bajo Delta del Río Paraná (Argentina).<br />

Darwiniana 41(1-4).<br />

K<strong>and</strong>us, P. 1997. Análisis de patrones de vegetación a escala regional en las islas del sector<br />

bonaerense del Delta de Río Paraná. Tesis doctoral, FCEN, Universidad de Buenos<br />

Aires.<br />

Minotti, P, Fernández, P <strong>and</strong> Corley, G., 1988. Regionalización del régimen de inundaciones en<br />

el Delta del Paraná; En: Adamoli, J. <strong>and</strong> A.I. Malvárez (eds). Condicionantes<br />

ambientales y bases para la formulacion de alternativas productivas en la región del<br />

Delta del río Paraná. Informe Final Proyecto UBACYT 135.<br />

Mujica, F., 1979. Estudio ecológico y socioeconómico del Delta Entrerriano. Parte I. Instituto<br />

Nacional de Tecnología Agropecuaria. Paraná. Argentina.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Fracassi & D. Somma 2010. Participatory action research concerning the l<strong>and</strong>scape use by a native cervid in a wetl<strong>and</strong><br />

231<br />

Petray, E. 2000. Las actividades relativas al cultivo y la utilización del álamo y del<br />

sauce. Período 1966-1999. Comisión Nacional del Álamo de Argentina.<br />

Pinder, L <strong>and</strong> Grosse. A., 1991. Blastoceros dichotomus. Mammalian Species. Nº 380 1-4.<br />

SAGPyA, 1999. Estado del avance del 1. Inventario de bosques cultivados a marzo de<br />

1999: Informe de Consultoria. Buenos Aires, 1999.<br />

Schaller, G.B. <strong>and</strong> Hamer, A., 1978. Rutting behabiour of Pére David´s deer, Elaphorus<br />

davidianus. Zool. Gart. N.F. 48, 1-15.<br />

Thornback J <strong>and</strong> Jenkins, M., 1982. The IUCN mammal red data book. Part 1: Threatened<br />

mammalian taxa of the Americas <strong>and</strong> the Austrasia zoogeographic region (excluding<br />

Cetacea). International Union Conservation Nature, Switzerl<strong>and</strong> 516 pp.<br />

Tomás, W., Beccaceci, M. <strong>and</strong> Pinder, L. 1997. Cervo-do-pantanal (Blastocerus dichotomus).<br />

En: Biologia e Conservaçao de Cervídeos Sul-americanos: Blastocerus, Ozotocerus e<br />

Mazama (Editor: Barbanti Duarte). FUNEP, Jaboticabal.<br />

Varela, D., 2003. Distribución, Abundancia y Conservación del Ciervo de los Pantanos<br />

(Blastocerus dichotomus) en el Bajo Delta del Río Paraná, Provincia de Buenos<br />

Aires, Argentina. Tesis de Licenciatura, Universidad de Buenos Aires (UBA),<br />

53pp.<br />

Wemmer, C., (Ed.), 1998. Deer. Status survey <strong>and</strong> conservation action plan. IUCN/SSC Deer<br />

Specialist Group. IUCN, Gl<strong>and</strong>, Suiza y Cambridge, UK. 106pp.<br />

Table 1. Principal threats for marsh deer in the Delta of the Paraná River according to interviewees.<br />

Threats<br />

Number of interviewees that<br />

listed this threat as the principal<br />

threat<br />

Number of interviewees that<br />

listed this threat as a secondary<br />

threat<br />

Hunting 16 3<br />

Habitat modification 8 4<br />

Interaction whit cattle 2 6<br />

Table 2. Perception of interviewees regarding the main actions contributing to the marsh deer<br />

conservation in the Delta of the Paraná River.<br />

Action Importance degree (%)<br />

Effective control of hunting 39.4<br />

Education & Awareness 27.3<br />

Maintenance patches of unmanaged understory 21.1<br />

afforestations<br />

Improve livestock good management practices 6.0<br />

Research on marsh deer ecology 3.1<br />

Hydrological management into afforestations 3.1<br />

Percentage of each Habitat<br />

Willow afforestation<br />

Poplar afforestation<br />

Afforestation without<br />

management<br />

Marshes<br />

Cattle pasture<br />

Riparian forest<br />

Roads<br />

Others<br />

Figure 1. Percentage of each habitat in the influence area of stakeholders.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Fracassi & D. Somma 2010. Participatory action research concerning the l<strong>and</strong>scape use by a native cervid in a wetl<strong>and</strong><br />

232<br />

Marsh deer habitat uses<br />

%<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Adult<br />

Willow<br />

Adult Poplar<br />

Young<br />

Poplar<br />

Young<br />

Willow<br />

Afforestation<br />

without<br />

management<br />

Marshes<br />

Cattle<br />

pasture<br />

Riparian<br />

forest<br />

roads<br />

others<br />

Habitat<br />

Figure 2. Percent of marsh deer records in each habitat typeas reported by interviewees<br />

Marsh deer Habitat Uses<br />

45<br />

40<br />

35<br />

30<br />

25<br />

%<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Adult Willow Adult Poplar<br />

Young<br />

Poplar<br />

Young<br />

Willow<br />

Afforestation<br />

without<br />

management<br />

Marshes<br />

Cattle<br />

pasture<br />

Riparian<br />

forest<br />

roads<br />

others<br />

Habitat<br />

Figure 3. Percent of interviewees that reported each habitat type as a refuge for the Marsh deer.<br />

Number of individuals in the Population<br />

70<br />

60<br />

50<br />

40<br />

%<br />

30<br />

20<br />

10<br />

0<br />

No change Fewer than 10 year ago Higher than 10 year ago<br />

Nº Individuals<br />

Figure 4. Perception of interviewees regarding the current number of marsh deers in the Delta population.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Hartter et al. 2010. Fortresses <strong>and</strong> fragments: impacts of fragmentation in a forest park l<strong>and</strong>scape<br />

233<br />

Fortresses <strong>and</strong> fragments: impacts of fragmentation in a forest park<br />

l<strong>and</strong>scape<br />

Joel Hartter 1* , Sadie J. Ryan 2 , Jane Southworth 3 & Colin A. Chapman4<br />

1 Department of Geography, University of New Hampshire, 102 Huddleston Hall, 73<br />

Main Street, Durham, NH 03825, USA<br />

2 National Center for Ecological Analysis <strong>and</strong> Synthesis (NCEAS), University of<br />

California, 735 State Street, Suite 300, Santa Barbara, CA 93101-5504, USA<br />

3 Department of Geography, L<strong>and</strong> Use <strong>and</strong> Environmental <strong>Change</strong> Institute (LUECI),<br />

University of Florida, 3141 Turlington Hall, Gainesville, FL 32611-7315, USA<br />

4<br />

Department of Anthropology <strong>and</strong> McGill School of Environment, McGill University,<br />

Montreal, Quebec, H3A 2T7, Canada<br />

Abstract<br />

Our research addresses patterns of l<strong>and</strong> cover <strong>and</strong> forest fragmentation in <strong>and</strong> around<br />

Kibale National Park in equatorial East Africa, <strong>and</strong> how park presence affects local<br />

livelihoods. Combining discrete <strong>and</strong> continuous data analyses of satellite imagery with<br />

a geographically r<strong>and</strong>om sample of two agricultural areas neighboring Kibale, we<br />

examine multi-scalar l<strong>and</strong>scape change <strong>and</strong> diminishing resources in the context of<br />

population increase, potential climate change, <strong>and</strong> fortress conservation. While park<br />

boundaries have remained relatively intact since 1984, the domesticated l<strong>and</strong>scape has<br />

become increasingly fragmented, with forests <strong>and</strong> wetl<strong>and</strong>s shrinking, becoming more<br />

isolated, <strong>and</strong> suffering decreased productivity. Remnant wetl<strong>and</strong> <strong>and</strong> forests are of<br />

particular interest because they supply ecological goods <strong>and</strong> services, but also provide<br />

habitat for primates <strong>and</strong> elephants from which to raid crops, not only posing a risk to<br />

food security, but may also lead to zoonotic disease emergence through spillover <strong>and</strong><br />

spillback events.<br />

Keywords: Kibale National Park, forest fragments, protected areas, l<strong>and</strong>scape fragmentation<br />

1. Introduction<br />

Recently, there has been a call for an integrated methodology that crosses temporal <strong>and</strong><br />

spatial scales to promote underst<strong>and</strong>ing of the social, ecological, <strong>and</strong> climatological<br />

dynamics within park l<strong>and</strong>scapes <strong>and</strong> to identify global trends, focus conservation<br />

priorities, <strong>and</strong> enable innovative <strong>and</strong> effective policy <strong>and</strong> resource management at<br />

multiple levels (DeFries et al. 2007). This paper details our data acquisition<br />

methodology that is designed to anticipate the consequences of a dynamic social <strong>and</strong><br />

ecological system faced with anthropogenic pressures <strong>and</strong> climate change at multiple<br />

scales to inform appropriate management or legislative interventions for the<br />

mechanisms using Kibale National Park as a “natural laboratory”.<br />

Establishing parks is the primary mechanism used to protect tropical forest<br />

biodiversity, particularly in regions with high human densities. Parks (protected areas of<br />

* Joel Hartter. Tel.: +1 603-862-7052 - Fax:+1 603-862-4362<br />

Email address: joel.hartter@unh.edu<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Hartter et al. 2010. Fortresses <strong>and</strong> fragments: impacts of fragmentation in a forest park l<strong>and</strong>scape<br />

234<br />

all sorts) protect <strong>and</strong> maintain endemic, threatened or endangered, flora <strong>and</strong> fauna,<br />

geological features, <strong>and</strong> cultural heritage sites. In addition, they can generate income<br />

for the local <strong>and</strong> national economies, <strong>and</strong> provide important benefits associated with<br />

enhanced tourism sectors. However, many parks are also associated with negative social<br />

<strong>and</strong> ecological impacts. Human populations neighboring parks are often excluded from<br />

settlement, access, resource extraction, <strong>and</strong> most forms of consumptive l<strong>and</strong> use in the<br />

park, which in turn affects farmers’ l<strong>and</strong> use <strong>and</strong> livelihood options.<br />

The processes that drive l<strong>and</strong> cover change are complex <strong>and</strong> cannot be understood<br />

without addressing underlying cause <strong>and</strong> effect relationships. <strong>Change</strong>s in climate,<br />

population, <strong>and</strong> l<strong>and</strong> use occur <strong>and</strong> interact simultaneously at different temporal <strong>and</strong><br />

spatial scales, having major implications for both livelihoods <strong>and</strong> biodiversity. <strong>Forest</strong><br />

loss <strong>and</strong> fragmentation are regarded as the greatest threat to global biological diversity<br />

(Turner <strong>and</strong> Corlett 1996). Fragmentation negatively impacts species composition due<br />

to a reduction in forest area <strong>and</strong> isolation of remaining fragments. Between 1990 <strong>and</strong><br />

2005, forest cover in Africa decreased by 21 million ha (Chapman et al. 2006). Since<br />

most parks are already ecosystem remnants of a limited size, it is important to consider<br />

each as a component of a larger l<strong>and</strong>scape.<br />

The presence <strong>and</strong> use of forest fragments outside parks illustrate the<br />

interconnected nature of the park-domestic l<strong>and</strong>scape.Within the greater continuous<br />

l<strong>and</strong>scape, remnant fragments of larger forests can be important to neighboring human<br />

communities, providing subsistence-based resources. These forests represent reservoirs<br />

of l<strong>and</strong>, resources, <strong>and</strong> economic opportunity for people, while at the same time are<br />

often viewed as buffers for parks by managers, as wildlife corridors <strong>and</strong> habitat that<br />

extends the effective size of the park.<br />

Unfortunately, these forest fragments are also hazards for local farmers, as sources<br />

of crop raids by primates, elephants, <strong>and</strong> birds. There has been extensive conversion of<br />

fragments, both to claim more l<strong>and</strong> <strong>and</strong> to destroy the habitat of would-be crop raiders.<br />

Health concerns have also arisen as pest animals move pathogens across l<strong>and</strong>scapes,<br />

leading to spillover of disease (potential zoonotic emergence), <strong>and</strong> spillback (pathogens<br />

transferred back into parks).<br />

The decline of remaining fragments in the human-dominated l<strong>and</strong>scape may be<br />

an inevitable process, exacerbated by the impacts of current <strong>and</strong> future changing climate.<br />

The impacts that fragmentation has on both wildlife <strong>and</strong> vegetation within a fragment<br />

<strong>and</strong> perhaps more importantly, the impact of loss of intact habitat <strong>and</strong> wildlife on the<br />

people relying on the remaining fragments, are important to underst<strong>and</strong>ing <strong>and</strong> slowing<br />

or preventing future decline. As fragments decrease <strong>and</strong> become more degraded,<br />

encroachment into the park <strong>and</strong> the number <strong>and</strong> severity of human-wildlife incidences<br />

may increase.<br />

There is growing interest in the academic community <strong>and</strong> by policy makers as to<br />

the extent of climate change impact on ecological communities (Walther et al. 2002).<br />

Climate variability <strong>and</strong> change are affecting wildlife populations, posing serious risk to<br />

poverty reduction <strong>and</strong> threatening rural livelihoods. Local climate variability directly<br />

impacts crop yields, resource quality <strong>and</strong> abundance, vegetation productivity, <strong>and</strong><br />

wildlife habitat <strong>and</strong> food sources. Climate change can also adversely affect the<br />

availability of resources for human populations outside of parks, who depend on the<br />

l<strong>and</strong> for their livelihoods. In addition, changes in food abundance within parks may<br />

force wildlife to seek food in agricultural areas near the park boundary, increasing the<br />

vulnerability of farms to crop damage <strong>and</strong> predation.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Hartter et al. 2010. Fortresses <strong>and</strong> fragments: impacts of fragmentation in a forest park l<strong>and</strong>scape<br />

235<br />

Figure 1. Kibale National Park in western Ug<strong>and</strong>a.<br />

2. Study Site<br />

Kibale National Park (795km 2 , Figure 1), medium altitude tropical moist forest in<br />

western Ug<strong>and</strong>a represents one of the best-studied forest sites in forested Africa, having<br />

been the site for multiple field projects for 40 years. The human population surrounding<br />

Kibale has increased seven-fold since 1920 <strong>and</strong> exceeds 270 people/km 2 at the western<br />

edge. Annual population growth rates range between 3 <strong>and</strong> 4% per year. The l<strong>and</strong>scape<br />

is a mosaic of small farms (most 200ha), <strong>and</strong> interspersed<br />

forest fragments <strong>and</strong> wetl<strong>and</strong>s (0.5-200ha or more), effectively isolating the park<br />

(Hartter <strong>and</strong> Southworth 2009). <strong>Forest</strong> fragments <strong>and</strong> wetl<strong>and</strong>s extending from Kibale’s<br />

boundaries <strong>and</strong> isolated within the agricultural matrix vary in size, shape, <strong>and</strong> resource<br />

types <strong>and</strong> amount. The Kibale l<strong>and</strong>scape is an extremely valuable setting because the<br />

biophysical <strong>and</strong> social response to social <strong>and</strong> ecological change may be emblematic of<br />

forest park l<strong>and</strong>scapes across the entire Albertine Rift <strong>and</strong> likely elsewhere. Parks in the<br />

Albertine Rift have been identified as areas of extremely high endemic biodiversity <strong>and</strong><br />

have been classed as a conservation priority (Plumptre et al. 2007).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Hartter et al. 2010. Fortresses <strong>and</strong> fragments: impacts of fragmentation in a forest park l<strong>and</strong>scape<br />

236<br />

3. Methodology<br />

Integrating l<strong>and</strong>scape level change detection with household level processes <strong>and</strong><br />

ecological sampling requires a sampling scheme that links household surveys <strong>and</strong><br />

ecological sampling to remotely sensed data. Social <strong>and</strong> ecological surveys provide<br />

valuable fine-scale, ground-data. Household-level decision making is critical to<br />

underst<strong>and</strong>ing the changes in l<strong>and</strong>-use <strong>and</strong> l<strong>and</strong>-cover <strong>and</strong> their effects on livelihoods.<br />

Ecological sampling provides means of underst<strong>and</strong>ing the consequences of household<br />

decisions to biodiversity <strong>and</strong> allows the generation of future scenarios for change.<br />

Linking classifications of l<strong>and</strong> cover to socio-economic <strong>and</strong> ecological surveys can<br />

provide a more comprehensive underst<strong>and</strong>ing of l<strong>and</strong> use <strong>and</strong> l<strong>and</strong> cover change over<br />

time. Instead of asserting that deforestation happened, we can now describe the effects<br />

of this change (i.e., loss of biodiversity), <strong>and</strong> pinpoint key social drivers for this change.<br />

Figure 2. Hierarchical framework that is embedded with multiple datasets at multiple spatial resolutions<br />

to examine four elements of the park l<strong>and</strong>scape <strong>and</strong> their temporal <strong>and</strong> spatial changes. Within this<br />

hierarchical framework, the higher level provides a context <strong>and</strong> imposes top-down constraints on the<br />

lower level, <strong>and</strong> the lower level provides mechanisms <strong>and</strong> imposes bottom-up constraints.<br />

3.1 Data Acquisition<br />

To provide the link between l<strong>and</strong>scape <strong>and</strong> household/ecological data, appropriate<br />

areas had to be defined that permitted the integration of research tool <strong>and</strong> intellectual<br />

questions. These areas had to be large enough to include the scales of satelliteevaluated<br />

l<strong>and</strong> cover change, <strong>and</strong> small enough to permit assessment of biodiversity <strong>and</strong><br />

human activities.<br />

Household Interviews – To link micro-scale l<strong>and</strong> use decisions with meso-scale l<strong>and</strong><br />

cover change area, we used the superpixel methodology (Hartter <strong>and</strong> Southworth 2009)<br />

to define a 5-km perimeter around park boundaries as the meso-scale research area.<br />

Interview respondents were selected from among l<strong>and</strong>holders in each of the 95 9-ha<br />

circular superpixels for which there were l<strong>and</strong>holders. The number of respondents<br />

selected per superpixel was proportional to the number of l<strong>and</strong>holders controlling l<strong>and</strong><br />

within the study area, <strong>and</strong> at least one interview was conducted in each superpixel.<br />

Therefore, superpixels with more l<strong>and</strong>holders (<strong>and</strong> correspondingly smaller individual<br />

l<strong>and</strong>holdings) had a higher sampling intensity than those with fewer l<strong>and</strong>holders. GPS<br />

points were also taken at access points for the nearest forest fragment <strong>and</strong> wetl<strong>and</strong> for<br />

each house <strong>and</strong> were used to calculate the straight-line distance from the house to the<br />

nearest wetl<strong>and</strong> <strong>and</strong>/or forest fragment <strong>and</strong> park boundary. Thus social <strong>and</strong> ecological<br />

units of interest can be mapped together.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Hartter et al. 2010. Fortresses <strong>and</strong> fragments: impacts of fragmentation in a forest park l<strong>and</strong>scape<br />

237<br />

Time Series Analysis of L<strong>and</strong> Cover, Productivity, & L<strong>and</strong>scape Fragmentation – To<br />

quantify l<strong>and</strong>scape change outside Kibale, L<strong>and</strong>sat TM <strong>and</strong> ETM+ imagery was chosen<br />

because it offers the best combination of spatial, spectral, temporal <strong>and</strong> radiometric<br />

resolutions. Six dry season images were acquired: (August 4, 1986, August 20, 1989,<br />

January 17, 1995, January 9, 2001, January 31, 2003, <strong>and</strong> September 1, 2008) (path<br />

162, row 60). An additional image (May 26, 1984) was acquired near the end of the<br />

rainy season <strong>and</strong> was the only available cloud-free image within this time period. Our<br />

analysis accounts for the phonological difference. Images were geometrically registered<br />

to 1:50,000 scale survey topographic maps of the region within an RMS of


J. Hartter et al. 2010. Fortresses <strong>and</strong> fragments: impacts of fragmentation in a forest park l<strong>and</strong>scape<br />

238<br />

over the period of record. Periodicities in autocorrelation functions are indicative of a<br />

cyclical response in seasonal <strong>and</strong> annual climate variables <strong>and</strong> may be attributed to<br />

fluctuations in large-scale atmospheric circulation patterns (e.g., ENSO). Spatial<br />

autocorrelation will be used to assess the spatial variability in temperature <strong>and</strong><br />

precipitation at sub-regional scales. High spatial autocorrelations between stations may<br />

be indicative of strong regional influences on local climate, whereas strong local<br />

influences, such as l<strong>and</strong> use, may result in lower spatial autocorrelations between<br />

stations.<br />

Integration of data – A series of logistic models will be incorporated into a model<br />

selection algorithm, assessed using an information theoretic framework, such as AIC<br />

(Akaike’s Information Criterion) based estimators. These models will be constructed<br />

by overlaying the information gathered <strong>and</strong> constructing process based models<br />

predicting fragment loss as a result of covariates such as ownership, fragment type,<br />

isolation <strong>and</strong> distance metrics, ecological factors, climate in preceding years,<br />

neighborly crop types <strong>and</strong> perceptions of use. We will derive suites of models at<br />

different spatial scales, wherein processes will likely differ.<br />

4. Conclusion<br />

While it is well established that models of the park-l<strong>and</strong>scape interface are necessary<br />

to the conservation discourse <strong>and</strong> to the persistence <strong>and</strong> sustainability of parks <strong>and</strong> the<br />

human populations that surround them, the necessary data to underst<strong>and</strong> the<br />

multifarious processes at play are hard to acquire. We have addressed issues of socioecological<br />

boundaries, climate <strong>and</strong> both spatial <strong>and</strong> temporal scale within our data<br />

acquisition. We will thus not only document a park-l<strong>and</strong>scape dynamic process, but<br />

identify drivers which are relevant to management <strong>and</strong> policy.<br />

References<br />

Chapman, C.A., Wasserman, M.D., Gillespie, T.R., Speirs, M.L., Lawes, M.J., Saj, T.L., <strong>and</strong><br />

T.E. Ziegler, 2006. Do nutrition, parasitism, <strong>and</strong> stress have synergistic effects on red<br />

colobus populations living in forest fragments American Journal of Physical<br />

Anthropology, 131: 525-534.<br />

DeFries, R., Hansen, A., turner, B.L., Redi, R., <strong>and</strong> J. Liu, 2007. L<strong>and</strong> use change around<br />

protected areas: management to balance human needs <strong>and</strong> ecological function. Ecological<br />

Applications, 17: 1031-1038.<br />

Hartter, J. <strong>and</strong> J. Southworth, 2009. Dwindling resources <strong>and</strong> fragmentation of l<strong>and</strong>scapes<br />

around parks: wetl<strong>and</strong>s <strong>and</strong> forest patches around Kibale National Park, Ug<strong>and</strong>a.<br />

L<strong>and</strong>scape Ecology, 24: 643-656.<br />

Stampone, M.D., Hartter, J., Ryan, S.J., <strong>and</strong> C.A. Chapman, Submitted. Temporal <strong>and</strong> spatial<br />

variability of rainfall in <strong>and</strong> around a forest park in western Ug<strong>and</strong>a. Journal of Applied<br />

Meteorology <strong>and</strong> Climatology.<br />

Plumptre, A.J., Behangana, M., Ndomba, E., Davenport, T., Kahindo, C., Kityo, R.<br />

Ssegawa, P., Eilu, G., Nkuutu, D. <strong>and</strong> I. Owiunji, 2003. The Biodiversity of the<br />

Albertine Rift. Wildlife Conservation Society, 107p.<br />

Turner, I. <strong>and</strong> R. Corlett, 1996. The conservation value of small, isolated fragments of lowl<strong>and</strong><br />

tropical rain forest. Trends in Ecology & Evolution 11: 330-333.<br />

Walther, G.R., et al., 2002. Ecological responses to recent climate change. Nature, 416: 389-395.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta et al. 2010. L<strong>and</strong>scape Structure of a Conservation Area <strong>and</strong> its Surroundings in Minas Gerais State, Brazil<br />

239<br />

L<strong>and</strong>scape structure of a conservation area <strong>and</strong> its surroundings in<br />

Minas Gerais State, Brazil<br />

Marise Barreiros Horta 1* , Carla Araújo Simões 2 , Eduardo Christófaro de Andrade 2 &<br />

Luciana Eler França 2<br />

1 Companhia Botânica, Brazil<br />

2 Delphi Projetos e Gestão Ltda., Brazil<br />

Abstract<br />

This study examined the suitability of a conservation area establishment through the l<strong>and</strong>scape<br />

structure investigation inside <strong>and</strong> outside the area. The l<strong>and</strong> cover mapping was performed<br />

using CBERS image, supervised <strong>and</strong> unsupervised classification. L<strong>and</strong>scape metrics were<br />

calculated utilizing Fragstats 3.3. The presence of a large forest patch inside the conservation<br />

area revealed the preservation of a rare patch in the l<strong>and</strong>scape. Shape index varied from 1,68 to<br />

2,02. Core index was higher for the forest patches (92,73%). Percentages of 50,00% of forest<br />

patches <strong>and</strong> 57,14% of grassl<strong>and</strong>s presented distances lower than 100m. The pastures showed<br />

high juxtaposition index (95,41%). The overall pattern denoted low influence of edge effect <strong>and</strong><br />

reasonable connectivity among patches. Conservation planning for the area has to take into<br />

consideration that grazing can be an impact source. Measures for the surrounding areas<br />

maintenance as a buffer zone can enhance flux among patches inside <strong>and</strong> outside the area.<br />

Keywords: L<strong>and</strong>scaspe structure, l<strong>and</strong>scape metrics, nature conservation<br />

1. Introduction<br />

Nature conservation in Brazil has been a challenge since the dem<strong>and</strong> for economic growth has<br />

contributed to environmental degradation, fragmentation <strong>and</strong> deforestation. The country<br />

presents a great importance for global biodiversity comprising a large percent of the remaining<br />

tropical forest stock. Although a large amount of l<strong>and</strong> is under protection these do not represent<br />

Brazil’s biodiversity. (Lele et al 2000).<br />

Biodiversity <strong>and</strong> ecological infrastructure maintenance are fundamental for nature conservation<br />

(Bridgewater 1993). The conservation value of a specific area in terms of ecological<br />

infrastructure can be examined through the measurement of the l<strong>and</strong>scape features (Haines-<br />

Young <strong>and</strong> Chopping 1996). In the case of biodiversity, the role of l<strong>and</strong>scape structure<br />

characterization is well known. Some l<strong>and</strong>scape components such as patch size, heterogeneity,<br />

perimeter-area ratio <strong>and</strong> connectivity can be major controls for species composition <strong>and</strong><br />

abundance (Noss <strong>and</strong> Harris 1986).<br />

As a part of the l<strong>and</strong>scape, the structure comprises the patches connected by corridors <strong>and</strong><br />

surrounded by a matrix (Forman <strong>and</strong> Godron 1986). The l<strong>and</strong>scape structure patterns include<br />

two aspects: composition <strong>and</strong> configuration (Turner 1989; Griffith et al 2000). L<strong>and</strong>scape<br />

composition refers to the patch or l<strong>and</strong> cover types without being spatially explicit. L<strong>and</strong>scape<br />

configuration has to do with the spatial distribution of patches or cover types within the<br />

* Corresponding author. Tel.: 55 31 33445937; 55 31 88835937<br />

Email address: marisehorta@companhiabotanica.com.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta et al. 2010. L<strong>and</strong>scape Structure of a Conservation Area <strong>and</strong> its Surroundings in Minas Gerais State, Brazil<br />

240<br />

l<strong>and</strong>scape <strong>and</strong> includes measures relatives to the placement of patch types relative to others<br />

(Mac Garigall <strong>and</strong> Marks 1995).<br />

A large number of l<strong>and</strong>scape indices that provide quantitative measurement of the l<strong>and</strong>scape<br />

structure are found in the literature (Forman <strong>and</strong> Godron 1986; Turner <strong>and</strong> Gardner 1990; Mac<br />

Garigall <strong>and</strong> Marks 1995; Haines-Young <strong>and</strong> Chopping 1996). According to Haines-Young <strong>and</strong><br />

Chopping (1996), for practical purposes, those indices can be grouped in the following<br />

categories: area, edge, shape, core area, nearest-neighbor, diversity, contagion <strong>and</strong> interspersion.<br />

The present study aimed to verify the l<strong>and</strong>scape structure concerning area, shape, core area,<br />

nearest-neighbor, contagion <strong>and</strong> interspersion inside <strong>and</strong> outside an area to be preserved. The<br />

conservation area is located in the Iron Quadrangle (Quadrilátero Ferrífero) region, Minas<br />

Gerais state, where besides the agricultural l<strong>and</strong> use pressure, the mineral resources exploitation<br />

has led to the reduction of rocky shrubl<strong>and</strong>s formations, grassl<strong>and</strong>s <strong>and</strong> forest. The<br />

responsibility for the conservation area creation belongs to the Vale Mining Company, as a<br />

dem<strong>and</strong> from the state forest government agency to compensate impacts on environmental<br />

resources, due to mining exploitation.<br />

2. Methodology<br />

2.1 Study area <strong>and</strong> l<strong>and</strong> cover mapping<br />

The conservation site Mata do Limoeiro is located in Itabira county, Minas Gerais state, in the<br />

west extreme distribution of the Brazilian Atlantic <strong>Forest</strong>, setting bounds with the Savannah<br />

dominion (Rizzini 1979). The climate is tropical showing mean annual temperature of 21,3 0 C.<br />

The relief is characterized by the presence of quartzite escarpments, gneiss hills <strong>and</strong> alluvial<br />

lowl<strong>and</strong>s (Radambrasil 1987).<br />

The l<strong>and</strong> cover mapping was performed for the conservation area <strong>and</strong> its surroundings (10 km<br />

round) using CBERS-2 CCD image 2008, 20m resolution. The software package Spring<br />

5.1.5/INPE was used for image processing. In a previous phase was carried out a contrast<br />

correction. Segmentation <strong>and</strong> classification techniques were applied during the processing phase.<br />

For the segmentation, a region growing method was used <strong>and</strong> for the classification, supervised<br />

(pixel based) <strong>and</strong> unsupervised (region based) procedures were undertaken. Adjustments in the<br />

classification were made after data collection in the field.<br />

2.2. L<strong>and</strong>scape pattern metrics<br />

The l<strong>and</strong>scape metrics were calculated utilizing the public domain software package Fragstats<br />

3.3. (Mac Garigal <strong>and</strong> Marks 1995). The l<strong>and</strong> cover map (raster format) was used as input data.<br />

The set of indices utilyzed for the l<strong>and</strong>scape structure characterization are listed in Table 1.<br />

Calculations of the relative percentage of the cover, number, size <strong>and</strong> isolation of the patches<br />

were also undertaken. The whole l<strong>and</strong>scape area was given by the sum of the area inside <strong>and</strong><br />

outside the conservation area.<br />

3. Results<br />

The l<strong>and</strong> cover map of the conservation area Mata do Limoeiro <strong>and</strong> its surroundings derived<br />

from the classification of the image CBERS-2 CCD 2008 is showed in the Figure 1. The l<strong>and</strong><br />

cover classes found were generalized into four: forest (semi-deciduous seasonal forest in<br />

advanced, intermediate <strong>and</strong> initial succession stages); grassl<strong>and</strong>s (savannah grassl<strong>and</strong>s, rocky<br />

shrubl<strong>and</strong>s, scrub); pasture (pasture, agricultural areas) <strong>and</strong> urban areas (villages, cities).<br />

The l<strong>and</strong>scape metrics resulted for the conservation site are presented in Table 1. The whole<br />

study area inside the conservation site is 2.097,04 ha. The forest class covers the largest area<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta et al. 2010. L<strong>and</strong>scape Structure of a Conservation Area <strong>and</strong> its Surroundings in Minas Gerais State, Brazil<br />

241<br />

(1.559,20ha), followed by the grassl<strong>and</strong>s (515,08ha) <strong>and</strong> pasture (31,76ha). When comparing<br />

the number of patches one can see that the majority belonged to the grassl<strong>and</strong>s comprising 35<br />

pieces, followed by 8 pasture <strong>and</strong> 6 forest pieces. The mean patch area was higher for forest<br />

class (258,36ha) that presented a continuous <strong>and</strong> large patch inside the conservation area of<br />

1.529,20ha. Patch density was 1,66 for grassl<strong>and</strong>s, 0,38 for pasture <strong>and</strong> 0,28 for forest.<br />

Table 1: L<strong>and</strong>scape metrics used for the l<strong>and</strong>scape structure characterization inside <strong>and</strong> outside the<br />

conservation area*<br />

Acronym<br />

Metric Name (units)<br />

Area Metrics<br />

TA<br />

Total Area (hectares)<br />

CA<br />

Total Class Area (hectares)<br />

PD<br />

Patch density (no./100 ha)<br />

NP<br />

Number of patches<br />

AREA MN<br />

Mean Patch Area (hectares)<br />

AREA<br />

Patch Size (patch size selected minimum <strong>and</strong> maximum)<br />

Shape Metrics<br />

SHAPE MN<br />

Mean shape index<br />

Core Metrics<br />

TCA Total Core Area (ha) – specified edge depth (20 m)<br />

CAI<br />

Core Area Index (percent)<br />

Isolation/Proximity Metrics<br />

ENN MN<br />

Mean euclidean nearest neighbor distance distribution (meters)<br />

Contagion/Interspersion Metrics<br />

IJI<br />

Interspersion <strong>and</strong> Juxtaposition Index (percent)<br />

* Full description of l<strong>and</strong>scape metrics equations is provided in Mac Garigal <strong>and</strong> Marks (1995)<br />

The mean shape index varied from 1,68 (grassl<strong>and</strong>s) to 2,02 (forest). The core area index was<br />

higher for forests. The mean distance among patches was lower for the grassl<strong>and</strong>s (109,82m),<br />

followed by pastures (164,83m) <strong>and</strong> forests (224,93m). The pastures showed high interspersion<br />

<strong>and</strong> juxtaposition index (95,41%). This index was 63,82% for grassl<strong>and</strong>s <strong>and</strong> 49,38% for the<br />

forest class.<br />

The results for the surrounding area are presented in Table 2. The total surrounding l<strong>and</strong>scape<br />

comprises an area of 52.674,92ha. The pastures cover the largest area (28.561,80ha), followed<br />

by forest (16.833,28 ha) <strong>and</strong> grassl<strong>and</strong>s (31,76ha). Two small urban areas are located inside the<br />

surrounding area occupying 35,56ha. The forest cover presented the highest number of patches<br />

(334), followed by pastures (224) <strong>and</strong> grassl<strong>and</strong>s (102). The mean patch area was higher for<br />

pastures (119,00ha). For the grassl<strong>and</strong>s <strong>and</strong> forest covers the value of this variable was<br />

respectively 71,02ha <strong>and</strong> 50,39ha. Patch density was 0,63 for forest, 0,45 for pasture <strong>and</strong> 0,19<br />

for forest.<br />

Mean shape index in the surrounding area varied from 1,69 (forest) to 2,10 (grassl<strong>and</strong>s). Core<br />

area index was similar for forest (86,58%) <strong>and</strong> grassl<strong>and</strong>s (85,91%). Mean distance among<br />

patches was lower for urban area (72,11m), followed by pasture (139,73m), forest (191,37m)<br />

<strong>and</strong> grassl<strong>and</strong>s (312,11m). The highest interspersion juxtaposition index was found for<br />

grassl<strong>and</strong>s (63,82%).<br />

Considering the whole l<strong>and</strong>scape area, the forest to be preserved inside the conservation area<br />

represents 8,43% <strong>and</strong> the grassl<strong>and</strong>s 6,63%. The 6 forest <strong>and</strong> 35 grassl<strong>and</strong>s patches were<br />

equivalent to 1,76% <strong>and</strong> 25,54% of the total patches. The largest forest patch inside the<br />

conservation area (> 1.000ha) represented 0,29% <strong>and</strong> the largest grassl<strong>and</strong>s one (> 100 ha)<br />

5,83%. Regarding isolation, 50,00% of forest patches <strong>and</strong> 57,14% of grassl<strong>and</strong>s presented<br />

distances lower than 100m.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta et al. 2010. L<strong>and</strong>scape Structure of a Conservation Area <strong>and</strong> its Surroundings in Minas Gerais State, Brazil<br />

242<br />

Figure 1: L<strong>and</strong> cover map of the conservation area Mata do Limoeiro <strong>and</strong> surroundings.<br />

Table 2: L<strong>and</strong>scape metrics results for the conservation area<br />

Metrics<br />

Classes<br />

<strong>Forest</strong> Grassl<strong>and</strong>s Pasture<br />

L<strong>and</strong>scape<br />

TA (ha) 1.550,20 515,08 31,76 2.097,04<br />

PD (no./100 ha) 0,28 1,66 0,38 2,33<br />

NP 6 35 8 49<br />

AREA MN (ha) 258,36 14,71 3,97 42,79<br />

AREA (minimum – ha) 0,60 0,04 0,04 0,04<br />

AREA (maximum – ha) 1.529,20 121,00 26,68 1.529,20<br />

SHAPE MN 2,02 1,68 1,73 1,73<br />

TCA (ha) 1.437,60 393,76 21,45 1.852,84<br />

CAI (%) 92,73 78,44 67,53 93,12<br />

ENN MN (m) 224,93 109,82 164,83 132,90<br />

IJI (%) 49,38 63,82 95,41 64,04<br />

Metrics<br />

Table 3: L<strong>and</strong>scape metrics results for the surrounding area<br />

Classes<br />

<strong>Forest</strong> Grassl<strong>and</strong>s Pasture Urban Areas L<strong>and</strong>scape<br />

TA ( ha) 16.833,28 7.244,28 28.561,80 35,56 52.674,92<br />

PD (no./100 ha) 0,63 0,19 0,45 0,00 1,28<br />

NP 334 102 240 2 678<br />

AREA MN (ha) 50,39 71,02 119,00 17,78 77,69<br />

AREA (min. – ha) 0,04 0,04 0,04 9,60 0,04<br />

AREA (max. – ha) 11.190,24 3.540,96 22.580,76 25,96 22.580,76<br />

SHAPE MN 1,69 2,10 1,78 1,88 1,77<br />

TCA (ha) 14.573,72 6.223,76 23.356,88 26,20 47.180,56<br />

CAI (%) 86,58 85,91 92,28 73,68 89,57<br />

ENN MN (m) 191,37 312,11 139,73 72,11 190,55<br />

IJI (%) 51,47 62,50 49,20 16,13 52,53<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta et al. 2010. L<strong>and</strong>scape Structure of a Conservation Area <strong>and</strong> its Surroundings in Minas Gerais State, Brazil<br />

243<br />

4. Discussion<br />

The l<strong>and</strong>scape structure results for the conservation site showed that the area to be preserved<br />

comprises a large amount of forest <strong>and</strong> grassl<strong>and</strong>s. These findings increase the site importance<br />

for nature conservation purposes especially considering that the area is set in the Brazil’s<br />

Atlantic <strong>Forest</strong> Dominion, a highly threatened tropical ecosystem (Conservational International<br />

2000), where conservation recommendations includes large areas, with elevate forest cover to<br />

favor species maintenance (Metzger 2009).<br />

Inside the conservation site the results of low patch number <strong>and</strong> high average area for the forest<br />

cover indicated the preservation of more continuous areas whereas in the surroundings the<br />

findings suggested concerns regarding fragmentation, especially for those areas located at the<br />

east side, closer to Itabira county. Valente (2001) found as more fragmented, in a l<strong>and</strong>scape<br />

structure study of river basins, the sites with lower mean patch area <strong>and</strong> higher patch density. As<br />

pointed out by Ji at al (2008), patch number <strong>and</strong> average area can reflect the fragmentation of a<br />

certain l<strong>and</strong>scape in some degree. <strong>L<strong>and</strong>scapes</strong> comprising lower mean patch area tend to be<br />

more fragmented (Mac Garigal <strong>and</strong> Marks 1995).<br />

The mean shape index results inside <strong>and</strong> outside the conservation area for the various l<strong>and</strong> cover<br />

classes indicated a prevalence of irregular patches. Valente (2001) <strong>and</strong> Jorge <strong>and</strong> Garcia (1997)<br />

found similar results for the mean shape index in forest <strong>and</strong> savannah environments. Patches<br />

characterized by larger areas tends to present more irregular shapes (Mac Garigal <strong>and</strong> Marks<br />

1995) <strong>and</strong> irregular patches with lower areas interact in a higher degree with the surrounding<br />

matrix being more susceptible to the edge effect (Valente 2001). In the present study, the<br />

characteristics of the core area might be compensating the shape. The l<strong>and</strong> cover majority inside<br />

<strong>and</strong> outside the conservation area presented high proportion of core area suggesting a larger<br />

proportion of natural vegetation with low influence of the edge effect. This is the case of the<br />

largest forest patch found inside the area to be preserved (1.529,20ha). Regarding conservation<br />

purposes, the forest patch exceeds the 250ha core area criterion suggested by Jongman (1995),<br />

for nature conservation planning in Europe.<br />

The mean distance among patches of the same l<strong>and</strong> cover inside the conservation area was<br />

highest for forests <strong>and</strong> in the surroundings for grassl<strong>and</strong>s. Considering that approximately half<br />

of the patches of natural vegetation inside the conservation area presented low mean distances, a<br />

relative aggregation <strong>and</strong> low isolation among patches is occurring. Low isolation enables a<br />

better influx of animals, seeds <strong>and</strong> pollen improving dispersion <strong>and</strong> biological diversity (Forman<br />

& Godron 1986).<br />

The highest interspersion <strong>and</strong> juxtaposition index inside the conservation area was found for the<br />

pastures. This l<strong>and</strong> cover type showed consequently a higher adjacency to the other patch types<br />

indicating a potential impact risk, since inappropriate fencing system coupled with free grazing,<br />

may favor invasive species colonization <strong>and</strong> damages to the seed bank.<br />

The surrounding area may work well as a buffer zone <strong>and</strong> enhance the flux among patches,<br />

given the richness in patches found. Nevertheless, the fragmentation trend especially on the east<br />

side <strong>and</strong> the large amount of pastures has to be considered. As pointed out by Jongman (1995)<br />

nature conservation sustainability must be supported by policies on buffering external<br />

influences <strong>and</strong> integrating functions, where possible.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta et al. 2010. L<strong>and</strong>scape Structure of a Conservation Area <strong>and</strong> its Surroundings in Minas Gerais State, Brazil<br />

244<br />

References<br />

Bridgewater, P.B., 1993. L<strong>and</strong>scape ecology, geographic information systems <strong>and</strong> nature<br />

conservation. In: Roy-Haines Young, David R. Green <strong>and</strong> Stephen H. Cousins (Eds.).<br />

L<strong>and</strong>scape Ecolgoy <strong>and</strong> GIS. Taylor & Francis, London, New York, Philadelphia: 23-36.<br />

Conservation International, 2000.Designing sustainable l<strong>and</strong>scapes. The Brazilian Atlantic<br />

Center for Applied Biodiversity Science. IESB.<br />

Forman, R.T.T. <strong>and</strong> Godron, M., 1986. L<strong>and</strong>scape Ecology. New York: John Wiley. 619 p.<br />

Griffith, G.A., Martinko, E.A. <strong>and</strong> Price, K.P., 2000. L<strong>and</strong>scape analysis of Kansas at three<br />

scales. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 52: 45-61.<br />

Haines-Young, R. <strong>and</strong> Chopping, M., 1996. Quantifying l<strong>and</strong>scape structure: a review of<br />

l<strong>and</strong>scape indices <strong>and</strong> their application to forested l<strong>and</strong>scapes. Progress in Physical<br />

Geography, 20: 418-445.<br />

Ji, J., Yuanyuan, C., Wunian, Y. <strong>and</strong> Yanian,K., 2008. Analysis <strong>and</strong> evaluation of l<strong>and</strong>scape<br />

pattern of Songpan county with remote sensing <strong>and</strong> GIS. The International Archives of<br />

the Photogrammetry, Remote Sensing <strong>and</strong> Spatial Information Sciences, 37: 1143-1148.<br />

Jongman, R.H.G., 1995. Nature conservation planning in Europe: developing ecological<br />

networks. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 32: 169-183.<br />

Jorge, L.A.B. <strong>and</strong> Garcia, G.J., 1997. A study of habitat fragmentation in southeastern Brazil<br />

using remote sensing <strong>and</strong> geographic information systems (GIS). <strong>Forest</strong> Ecology <strong>and</strong><br />

Management, 98: 35-47.<br />

Lele, U., Vianna, V., Verissimo, A., Vosti, S., Perkins, K. <strong>and</strong> Husain, S.A., 2000. Brazil.<br />

<strong>Forest</strong>s in balance: Challenges of conservation with development. Evaluation Country<br />

Case Studies Series. The World Bank. Washington D.C.<br />

Mac Garigal, K. <strong>and</strong> Marks, B.J., 1995. Fragstats: Spatial Pattern Analysis Program for<br />

Quantifying L<strong>and</strong>scape Structure. Gen.Tech.Report.US Department of<br />

Agriculture.Pacific Northwest Research Station. 122p.<br />

Metzger, J.P., 2009. Conservation issues in the Brazilian Atlantic <strong>Forest</strong>. Conservation Biology,<br />

142:1138-1140.<br />

Noss, R.F. <strong>and</strong> Harris, L.D., 1986.Nodes, networks <strong>and</strong> MUMs: preserving diversity at all<br />

scales. Environmental Management, 10: 299-209.<br />

Radambrasil, 1987. Geologia, geomorfologia, pedologia, vegetação e uso potencial da terra.<br />

Folha SE 24, Rio Doce.<br />

Rizzini, C. T., 1979. Tratado de Fitogeografia do Brasil. Aspectos sociológicos e florísticos.<br />

( Vol. 2). São Paulo: Hucitec - Edusp.<br />

Turner, M.G., 1989. L<strong>and</strong>scape ecology: the effect of pattern on process. Annual Review of<br />

Ecology <strong>and</strong> Systematics, 20:191-197.<br />

Turner, M.G. <strong>and</strong> Gardner, R.H., 1990. Quantitative methods in L<strong>and</strong>scape Ecology: An<br />

Introduction. In: Monica G. Turner <strong>and</strong> Robert H. Gardner (Eds.). Quantitative methods<br />

in L<strong>and</strong>scape Ecology: the analysis <strong>and</strong> interpretation of l<strong>and</strong>scape<br />

heterogeneity.Ecological Studies, Vol. 82. Springer-Verlag, New York: 3-14.<br />

Turner, M.G., 2005. L<strong>and</strong>scape Ecology: what is the state of the science.<br />

Annu.Rev.Ecol.Evol.Syst.36:319-344.<br />

Valente, R.O.A., 2001. Análise da estrutura da paisagem no rio Corumbataí, SP.<br />

Unpublished Dissertação de Mestrado, Universidade de São Paulo, Piracicaba, SP.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.O. López-Martínez et al. 2010. Ecological factors influencing beta diversity at two spatial scales in a tropical dry forest<br />

245<br />

Ecological factors influencing beta diversity at two spatial scales in a<br />

tropical dry forest of the Yucatán Peninsula<br />

J. Omar López-Martínez * , J. Luis Hernández-Stefanoni & Juan Manuel Dupuy<br />

Centro de Investigación Científica de Yucatán A.C. Unidad de Recursos Naturales,<br />

Calle 43 # 130. Colonia Chuburná de Hidalgo. C.P. 97200, Mérida, Yucatán, México<br />

Abstract<br />

Underst<strong>and</strong>ing the ecological factors determining beta diversity at different spatial<br />

scales is relevant for ecological theory <strong>and</strong> for conservation <strong>and</strong> management. We analyzed the<br />

relationship of beta-diversity <strong>and</strong> environmental <strong>and</strong> spatial variables at two spatial scales. We<br />

identified vegetation classes based on a supervised classification, determined species<br />

composition, <strong>and</strong> obtained soil samples from a total of 276 sites in 23 sampling l<strong>and</strong>scapes.<br />

Using vegetation classes <strong>and</strong> FRAGSTAT, we calculated patch-type metrics in each l<strong>and</strong>scape.<br />

Spatial dependence was included in the models using PCNM. Partial CCA <strong>and</strong> RDA were<br />

performed to partition variation. Soil properties, st<strong>and</strong>-age, l<strong>and</strong>scape metrics <strong>and</strong> site PCNMs<br />

were used at the local level, while l<strong>and</strong>scape metrics <strong>and</strong> l<strong>and</strong>scape PCNMs were used at the<br />

l<strong>and</strong>scape level. At the local level, space was the most important variable related to variation in<br />

species composition (48.87%), whereas at the l<strong>and</strong>scape-level, l<strong>and</strong>scape-metrics explained<br />

most (50%) of the variation in species composition.<br />

Keywords: Beta diversity; CCA; forest succession; l<strong>and</strong>scape patterns; RDA; spatial<br />

dependence<br />

1. Introduction<br />

Beta diversity refers to variation in species composition among sites within a<br />

geographic area (Legendre et al. 2005). It could be determined by different habitats along an<br />

environmental gradient, or by geographic distance within seemingly uniform habitats. In the<br />

latter case, beta diversity reflects spatial isolation among species (Halfter 1998). Beta diversity<br />

allows us to underst<strong>and</strong> how species habitats are distributed, it is also an important component<br />

in biodiversity conservation planning <strong>and</strong> management (Halfter 1998).<br />

Factors affecting species turnover <strong>and</strong> the scale at which they operate have been debated<br />

for a long time in ecology. Two of the main causal factors of beta diversity are: 1) limitation in<br />

the dispersal capacity of species that are demographically <strong>and</strong> competitively equal (neutral<br />

theory, Hubbell 2001). 2) Environmental conditions as a determinant of species presence <strong>and</strong>/or<br />

abundance (niche theory, Grubb 1977).On the other h<strong>and</strong>, it is important to note that factors<br />

affecting species turnover may vary among spatial scales (Garcia 2006).<br />

The aim of this study was to analyze the relative importance of environmental<br />

heterogeneity, st<strong>and</strong> age, l<strong>and</strong>scape structure <strong>and</strong> spatial dependence on species turnover of a<br />

tropical dry forest at two different spatial scales.<br />

* Corresponding author.<br />

Email address: jmartinez.omar@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.O. López-Martínez et al. 2010. Ecological factors influencing beta diversity at two spatial scales in a tropical dry forest<br />

246<br />

2. Methods<br />

2.1 Study Area<br />

The study was conducted in a l<strong>and</strong>scape of 22 x 16 km 2 located in the Yucatán<br />

Peninsula, México (-89.60 W, 20.01 N <strong>and</strong> -89.39 W, 20.16 N). The climate is warm<br />

subhumid, with summer rain (May-October) <strong>and</strong> a marked dry season (November-April). Mean<br />

annual temperature is ± 26°C, <strong>and</strong> mean annual precipitation ranges from 1000 to 1200 mm.<br />

Topography consists of flat areas alternating with low hills. Elevation ranges between 60 <strong>and</strong><br />

160 m.a.s.l (Flores <strong>and</strong> Espejel, 1994). The l<strong>and</strong>scape is dominated by tropical semideciduous<br />

forest of different successional ages derived from traditional agriculture (Figure 1).<br />

2.2 Remotely sensed data, imagery processing <strong>and</strong> calculation of l<strong>and</strong>scape metrics<br />

A Spot 5 satellite imagery acquired in January 2005 was geo-referenced to<br />

Universal Transverse Mercator Projection (WGS 84) <strong>and</strong> radiometrically corrected to minimize<br />

the effect of atmospheric scattering. A false color composite image was created from b<strong>and</strong>s 2<br />

(red), 3 (near infrared), <strong>and</strong> 4 (mid infrared). Training sites were selected from this image to<br />

perform the classification. The l<strong>and</strong>-cover classes consdiered were: 1) 3-8 y-old secondary<br />

forest; 2) 9-15 y-old secondary forest; 3) >15 y-old secondary forest on flat areas; 4) >15 y-old<br />

secondary forest on hills; 5) agricultural fields; 6) urban areas <strong>and</strong> roads. We used the maximum<br />

likelihood algorithm in Idrisi Kilimanjaro (Eastman, 2004), <strong>and</strong> two accuracy measures were<br />

applied to the map: the overall accuracy, <strong>and</strong> Cohen’s Kappa statistic (Campbell, 1987).<br />

We selected three l<strong>and</strong>scape metrics that can be considered as fragmentation indices:<br />

TECI (Total Edge Contrast Index), SIEI (Simpson´s Evenness Index) <strong>and</strong> LPI (Large Patch<br />

Index). The definition of these metrics is given in McGarigal et al (2002). Finally, the weighted<br />

edge contrast between vegetation cover classes required to compute total edge contrast index<br />

were calculated as the inverse values of the Morisita-Horn measures between each pair of<br />

vegetation cover classes.<br />

2.3 Species composition <strong>and</strong> environmental data.<br />

Survey sampling was performed in the summer of 2008 <strong>and</strong> 2009 during the rainy season. First,<br />

we selected 23 1 km 2 l<strong>and</strong>scapes along a fragmentation gradient. Second, we selected 12 sites<br />

using a stratified r<strong>and</strong>om sampling design using the four vegetation classes. Each site consisted<br />

of two concentric circular plots: all woody plants > 5 cm in DBH (diameter at breast (1.3 m)<br />

height), hereafter referred to as adults, were sampled in a 200 m 2 plot; whereas 1-5 cm DBH<br />

woody plants, hereafter referred to as juveniles, were sampled in a nested 50 m 2 subplot. In each<br />

site, we identified each individual <strong>and</strong> calculated the Importance Value Index (IVI) for each<br />

species. To characterize soil conditions at each site, we estimated percent stoniness <strong>and</strong><br />

collected soil samples for physical-chemical analyses: pH in water, electric conductivity<br />

(EC), soil organic matter (SOM, combustion or Walkley-Black), total nitrogen (N,<br />

Kjeldahl), available phosphorous (P, Bray), <strong>and</strong> interchangeable potassium (K).<br />

2.4 Statistical Analysis<br />

To explore general patterns of species composition, we used detrended<br />

correspondence analysis (DCA) ordination of the IVI of each species in each site or<br />

l<strong>and</strong>scape in CANOCO (4.52).<br />

Principal coordinate of neighbor matrices (PCNM) was used to account for spatial<br />

dependence (Borcard <strong>and</strong> Legendre 2002). The PCNM vectors were calculated from the<br />

location of each sampling site for local-level analyses <strong>and</strong> from the centroid of each l<strong>and</strong>scape<br />

unit for l<strong>and</strong>scape-level analyses.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.O. López-Martínez et al. 2010. Ecological factors influencing beta diversity at two spatial scales in a tropical dry forest<br />

247<br />

Variation in community composition among sites was partitioned using the method of<br />

Borcard et al (1992). Four sets of explanatory variables were considered at the local level:<br />

environmental variables (soil attributes, % stoniness), l<strong>and</strong>scape structure (TECI, SIEI, LPI),<br />

spatial dependence (PCNMs), <strong>and</strong> st<strong>and</strong> age; whereas only two sets of variables were used at<br />

the l<strong>and</strong>scape level: l<strong>and</strong>scape structure <strong>and</strong> spatial dependence. To determine the amount of<br />

variation explained independently by each set of predictor variables, we performed partial<br />

constrained ordination separately for each set of variables at both levels. Due to the length of<br />

gradients, CCA was used at the local level (3.99 SD), while RDA was used at the l<strong>and</strong>scape<br />

scale (2 SD).<br />

3. Results<br />

The l<strong>and</strong> cover thematic map of the study area is shown in Figure 1. The total area<br />

occupied by this l<strong>and</strong>scape was 37,243 ha, 94.8% corresponded to tropical sub-deciduous forest<br />

in any of the four vegetation classes, whereas 5.2% corresponded to agriculture <strong>and</strong> urban areas.<br />

We obtained an overall accuracy in the supervised classification of 73.3% <strong>and</strong> the Kappa index<br />

was of 0.7.<br />

Figure 1: Location <strong>and</strong> l<strong>and</strong> cover map of the study area obtained from a supervised classification.<br />

A total of 22,262 woody individuals belonging to 204 species <strong>and</strong> 52 families were<br />

recorded in 276 sites in 23 l<strong>and</strong>scapes. Of total of sites sampled, 51 belonged at class 1, 77 to<br />

class 2, 83 to class 3 <strong>and</strong> 62 to class 4. The most abundance species was Neomillspaughia<br />

emarginata with 3,421 individuals, whereas 29 species were represented only by a single<br />

individual.<br />

3.1 Local scale.<br />

Detrended correspondence analysis at the local scale showed a large length gradient of species<br />

composition (3.99 sd; Fig. 2a). Vegetation class 4 differed significantly from the other classes.<br />

Canonical correlation analysis showed that total variation explained was 13.79%. Variation<br />

partitioning showed that space is the most important variable explaining 5.95% of species<br />

composition, followed of the soil variables with 5.83% (Table 1). Vegetation class 1 was<br />

negatively associated with organic matter (SOM), stoniness, st<strong>and</strong> age <strong>and</strong> TECI whereas<br />

vegetation class 4 showed the opposite pattern (Fig. 2b).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.O. López-Martínez et al. 2010. Ecological factors influencing beta diversity at two spatial scales in a tropical dry forest<br />

248<br />

Figure 2. Ordination analysis at the local scale. (a) Detrended correspondence analysis (DCA) of the IVI<br />

of each species at each site. (b) Canonical correspondence analysis biplot for the first <strong>and</strong> second axes<br />

showing st<strong>and</strong> age, space, environmental, <strong>and</strong> l<strong>and</strong>scape structure variables.<br />

Table 1. Variation partitioning of woody plant species composition at the local scale.<br />

PREDICTOR<br />

VARIABLE<br />

MARGINAL<br />

VARIATION<br />

EXPLAINED<br />

PERCENT<br />

VARIATION<br />

EXPLAINED<br />

Soil Variables 0.49 41.56<br />

St<strong>and</strong> age 0.07 6.34<br />

Structure Metrics 0.08 6.94<br />

Space 0.50 43.10<br />

Share 0.02 2.06<br />

Total inertia explained 1.17 100<br />

Unknowledge 7.23<br />

Total Inertia 8.40<br />

3.2 L<strong>and</strong>scape scale.<br />

Detrended correspondence analysis showed a short length gradient (2 sd) of species<br />

composition. We found no clear pattern of species composition in relation to fragmentation<br />

class (Fig. 3a) (. Redundancy analysis showed that l<strong>and</strong>scape structure variables explained a<br />

greater amount of variation in species composition than space (16.5% <strong>and</strong> 13.8%, respectively;<br />

Table 2). Fragmentation class 4 was positively associated with SHEI <strong>and</strong> negatively associated<br />

with LPI, whereas fragmentation class 1 showed the opposite pattern (Fig. 3b).<br />

Figure 3. Ordination analysis at the local scale. (a) Principal components analysis (PCA) ordination of the<br />

IVI of each species in each site or l<strong>and</strong>scape. (b) Redundancy analysis biplot for the first <strong>and</strong> third axes<br />

showing Shannon´s Evenness Index (SHEI) <strong>and</strong> Large Patch Index variables.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.O. López-Martínez et al. 2010. Ecological factors influencing beta diversity at two spatial scales in a tropical dry forest<br />

249<br />

Table 2. Variation partitioning of woody plant species composition at the l<strong>and</strong>scape scale.<br />

4. Discussion<br />

MARGINAL PERCENT<br />

PREDICTOR<br />

VARIABLE VARIATION VARIATION<br />

EXPLAINED EXPLAINED<br />

L<strong>and</strong>scape structure 0.17 16.5<br />

Space 0.14 13.8<br />

Shared 0.01 1.4<br />

Total inertia explained 0.32 31.7<br />

Unknown 0.68 68.3<br />

Total Inertia 1.00<br />

The amount of variation in species composition (beta diversity) was higher at the<br />

l<strong>and</strong>scape scale (32%), than at the local scale (13.89%). This indicates that differences in<br />

species composition are greater among l<strong>and</strong>scape units with varying degrees of fragmentation,<br />

compared to differences among vegetation cover classes (Figures 2a, 3a).<br />

Arroyo-Mora et al. (2005) also found differences in species composition among<br />

successional classes in neotropical dry forests.<br />

At both spatial scales, environmental factors <strong>and</strong> spatial dependence explained a similar<br />

amount of variation in species composition (Tables 1. 2). This lends support to both niche<br />

theory (environmental drivers, Grubb, 1977) <strong>and</strong> the neutral theory (based on dispersal<br />

limitation, Hubbell 2001, Chust et al. 2006). This is consistent with the recently proposed<br />

continuum hypothesis of Gravel et al. (2006), which states that both niche theory <strong>and</strong> dispersal<br />

limitation play an important role in beta diversity.<br />

Variation partition at the local scale showed that soil attributes had a higher contribution<br />

to explaining beta diversity than st<strong>and</strong> age or l<strong>and</strong>scape structure (Table 1). In particular, soil<br />

organic matter <strong>and</strong> pH were the variables most strongly associated with change in species<br />

composition (Figure 2b). At the l<strong>and</strong>scape scale, l<strong>and</strong>scape structure was the most important<br />

contributor to beta diversity (Table 2). The large patch index (LPI), which was the variable most<br />

strongly associated with beta diversity, was also positively associated with the least disturbed<br />

l<strong>and</strong>scape units (Figure 3b). These results concur with other studies indicating that species<br />

composition is affected by several factors <strong>and</strong> process operating at different spatial scales<br />

(Leibold, 2006, Hubbell, 2001, Parker, 2004).<br />

Although there were fewer environmental <strong>and</strong> space variables explaining species<br />

composition at the l<strong>and</strong>scape scale than at the local scale, the variation explained by these<br />

variables was more than double at the l<strong>and</strong>scape scale compared to the local scale (31.7% versus<br />

13.8%). This may suggest that disturbance has a greater influence on woody species<br />

composition than vegetation cover.<br />

References<br />

Arroyo-Mora J.P., Sánchez-Azofeifa G.A., Kalácska M., Rivard B. <strong>and</strong> Janzen D.H. 2005.<br />

Secondary forest detection in a Neotropical dry forest using L<strong>and</strong>sat 7 ETM+ imagery.<br />

Biotropica, 37 (4): 497-507.<br />

Borcard, D. <strong>and</strong> Legendre, P., 2002. All-scale spatial analysis of ecological data by means of<br />

principal coordinates of neighbour matrices. Ecological Modelling, 153: 51-68.<br />

Borcard, D., Legendre, P. Drapeau, P., 1992. Partialling out the Spatial Component of<br />

Ecological Variation. Ecology, 73(3): 1045-1055.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.O. López-Martínez et al. 2010. Ecological factors influencing beta diversity at two spatial scales in a tropical dry forest<br />

250<br />

Chust, G., Chave, J., et al. 2006. Determinants <strong>and</strong> spatial modeling of tree β-diversity in a<br />

tropical forest l<strong>and</strong>scape in Panama. Journal of Vegetation Science, 17: 83-92.<br />

Eastman, R. 1987-2004. Idrisi (kilimanjaro) 14.02. Clark Labs, Clark University 950 Main<br />

Street, Worcester MA.<br />

Flores, S. & Espejel, I., 1994. Tipos de vegetación de la Península de Yucatán. Fascículo de la<br />

Etnoflora Yucatanense. Univ. Aut. de Yucatán. UADY.<br />

García, D., 2006. La escala y su importancia en el análisis espacial." ecosistemas. Revista<br />

Científica de Ecología y Medio Ambiente, 15(3): 7-18.<br />

Gravel, D., Canham, D. et al, 2006. Reconciling niche <strong>and</strong> neutrality: the continuum hypothesis.<br />

Ecology Letters, 9(399-409).<br />

Grubb, P. J., 1977. The maintenance of species-richness in plant communities: the importance<br />

of the regeneration niche. Biological reviews, 52(1): 107-145.<br />

Hubbell, S.P. (2001). The Unified Neutral Theory of Biodiversity <strong>and</strong> Biogeography. Princeton<br />

University Press.<br />

Legendre, P., D. Borcard, Peres-Neto, P. R., 200. Analyzing beta diversity: partitioning the<br />

spatial variation of community composition data. Ecological Monographs, 75(4): 435-<br />

450.<br />

Leibold, M. A. <strong>and</strong> McPeek, M. A., 2006. Coexistence of the niche <strong>and</strong> neutral perspectives in<br />

community ecology. Ecology, 87(6): 1399-1410.<br />

McGarigal, K., Cushman, S.A., Neel, M.C., Ene, E., 2002. FRAGSTATS: Spatial Pattern<br />

Analysis Program for Categorical Maps. Computer software program produced by the<br />

authors at the University of Massachusetts, Amherst available at the following web site:<br />

http://www.umass.edu/l<strong>and</strong>eco/research/fragstats/fragstats.html.<br />

Parker, V. T., 2004. The community of an individual: implications for the community concept.<br />

OIKOS, 104: 27-34.<br />

Halffte, G., 1998. Una estrategia para medir la biodiversidad a nivel de pasiaje. In: Halffter, G<br />

(Ed.). La diversidad biológica de Iberoamérica Vol. II. Acta Zoologica Mexicana, nueva<br />

serie. Volumen especial: 3-17.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Sara Marques et al. 2010. Impact of roads on ungulate species: a preliminary approach in Portugal<br />

251<br />

Impact of roads on ungulate species: a preliminary approach in<br />

Portugal<br />

Sara Marques 1,* , Catarina Ferreira 1 , Rogério Rodrigues 2 , Jorge Cancela 3 & Carlos<br />

Fonseca 1<br />

1 Biology Department & C<strong>ESA</strong>M, University of Aveiro, 3810-193 Aveiro, Portugal<br />

2 Northern Regional Direction of <strong>Forest</strong>s, National <strong>Forest</strong>ry Authority, 5000-567 Vila Real,<br />

Portugal<br />

3 Central Regional Direction of <strong>Forest</strong>s, National <strong>Forest</strong>ry Authority, 3500-093 Viseu,<br />

Portugal<br />

Abstract<br />

The impact of linear infrastructures, particularly the roads on wildlife, has been one of the<br />

highlight themes of a new field of knowledge that combine Conservation Biology <strong>and</strong> modern<br />

Engineering Sciences. The impacts of roads in the ecological l<strong>and</strong>scape include habitat loss,<br />

fragmentation <strong>and</strong> degradation, barrier effect, wildlife vehicle-collisions, among others longterm<br />

effects. Wildlife vehicle-collisions represent an additive source of mortality to wildlife<br />

populations in addiction to natural mortality, such as predation <strong>and</strong> disease. This preliminary<br />

approach reveals the importance of variables involved in vehicle collisions with ungulate<br />

species in northern <strong>and</strong> central Portugal, such as road type, time of the day <strong>and</strong> season of the<br />

year. In this study the majority of accidents occured with wild boar at national roads (EN),<br />

during the night <strong>and</strong> principally in winter <strong>and</strong> August. These studies are very important in order<br />

to prevent <strong>and</strong> mitigate effects of vehicle-collisions on large animal populations.<br />

Keywords: ungulate vehicle-collisions; wild boar; red deer; roe deer; central <strong>and</strong> northern<br />

Portugal<br />

1. Introduction<br />

The references of wildlife vehicle-collisions began on the 1920’ <strong>and</strong> 30’ by some researchers,<br />

<strong>and</strong> very few of these early studies found ungulate causalities (Gagnon et al. 2007). From 1970’<br />

big game species, such as ungulates, became a concern as traffic levels <strong>and</strong> vehicle speeds<br />

increased, leading to higher rates of ungulate vehicle-collisions. Research of potential direct <strong>and</strong><br />

indirect effects of roads <strong>and</strong> traffic on ungulates populations start on that time. There are many<br />

studies about this subject (Christie & Nason 2003, Seiler 2004, Huijser et al. 2009, Apollonio et<br />

al. 2010). Actually we know that roads <strong>and</strong> traffic associated have many primary effects on wild<br />

ungulates populations <strong>and</strong> ecological l<strong>and</strong>scape, as mortality caused by ungulate-vehicle<br />

collisions, habitat loss, fragmentation, <strong>and</strong> degradation, decreased movement across roadways<br />

leading to habitat fragmentation <strong>and</strong> potentially genetic isolation, recognized as barrier effect,<br />

among others long-term effects (Gagnon et al. 2007; Huijser et al. 2009).<br />

Roads, in general, are a great linear feature within the l<strong>and</strong>scape directly impacting wildlife<br />

populations through vehicle collisions. The ungulate vehicle-collisions often result in injuries or<br />

* Corresponding author.<br />

Email addresse: sara.marques@ua.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Sara Marques et al. 2010. Impact of roads on ungulate species: a preliminary approach in Portugal<br />

252<br />

fatalities to vehicle occupants, significant property damage, <strong>and</strong> animal deaths that represent an<br />

additive source of mortality to wildlife populations, in addition to other mortality, such as<br />

predation <strong>and</strong> disease (Christie <strong>and</strong> Nason 2004, Seiler 2004). Road mortality affects the<br />

individual animals <strong>and</strong> also some species at the population level which may create a serious<br />

reduction in population survival probability, affecting the entire ecosystem (Bruinderink <strong>and</strong><br />

Hazebroek 1996; Huijser 2009). Further, these collisions are a considerable threat to traffic<br />

safety, socio-economics, animal welfare, <strong>and</strong> wildlife management <strong>and</strong> conservation (Christie<br />

<strong>and</strong> Nason 2004; Seiler 2004; Huijser 2009).<br />

Some researchers state that the increasing traffic levels are the primary reason for number grow<br />

of ungulate vehicle-collisions but many recognize traffic level as a component of this increase,<br />

along others factors such as wildlife population fluctuations, wildlife behavior, driver behavior<br />

<strong>and</strong> temporal <strong>and</strong> spatial environmental factors (Gagnon et al. 2007).<br />

In Portugal there is a lack of information about this subject but we know that the wild ungulate<br />

populations, as wild boar (Sus scrofa), red deer (Cervus elaphus) <strong>and</strong> roe deer (Capreolus<br />

capreolus) tend to increase in a generalized way (Vingada et al. 2010). Some populations have<br />

been dispersing into historical areas like coastal <strong>and</strong> urban areas. It’s just at this range, where<br />

human presence is more intense, causing impacts on these species.<br />

This study aimed to put in evidence some variables involved in car accidents with ungulates in<br />

central <strong>and</strong> northern Portugal, <strong>and</strong> the conclusions obtained may be important to help the<br />

mitigation of ungulate-vehicle collisions <strong>and</strong> ungulate habitat fragmentation because it could<br />

potentially be used by roads designers <strong>and</strong> planners to avoid potentially hazardous areas in<br />

future roads development or new roadway design or to identify appropriate preventive measures.<br />

The findings could also potentially be used to identify “black points” on existing routes that are<br />

involved in many ungulate-vehicle accidents <strong>and</strong> should be the focus of mitigate procedures.<br />

2. Methodology<br />

Ungulate vehicle-collisions data was acquired from the National <strong>Forest</strong>ry Authority (AFN) of<br />

the Portuguese Ministry of Agriculture, particularly the services at central <strong>and</strong> northern Portugal<br />

(Central Regional Direction of <strong>Forest</strong>s (DRFC) <strong>and</strong> Northern Regional Direction of <strong>Forest</strong>s<br />

(DRFN)). Many of this data belong to the National Policy whose reports wildlife vehiclecollisions<br />

cases to AFN.<br />

Data used for this work include the districts of Aveiro, Braga, Bragança, Castelo Branco,<br />

Coimbra, Guarda, Leiria, Portalegre, Porto, Viana do Castelo, Vila Real <strong>and</strong> Viseu. The<br />

information was collected during 10 years, from 1999 to 2009. However data from 1999 <strong>and</strong><br />

2000 are unreliable, as well as in 2008 <strong>and</strong> 2009. During these two last years, the lack of<br />

information can be explain by the few number of cases that arrives to AFN. Probably some of<br />

them will be received soon.<br />

A database with important information as the species involved in the accident, local, the road<br />

type, date, particularly the month <strong>and</strong> year, <strong>and</strong> time of day was created. This study involves<br />

three ungulate species, wild boar (Sus scrofa), red deer (Cervus elaphus) <strong>and</strong> roe deer<br />

(Capreolus capreolus) <strong>and</strong> are considered six different road types namely national roads (EN),<br />

municipal roads (EM), high speed roads (A), complementary roads (IC), main roads (IP) <strong>and</strong><br />

forestry ways (CF). The times of day (dawn, day, evening <strong>and</strong> night) was defined according to<br />

the seasons with specific hour range.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Sara Marques et al. 2010. Impact of roads on ungulate species: a preliminary approach in Portugal<br />

253<br />

The information contained in database was subjected of a basic statistical analysis as the<br />

collisions’ percentage with wild boar, roe deer <strong>and</strong> roe deer, to underst<strong>and</strong> which the ungulate’s<br />

species are more affected by the vehicle accidents; collisions’ percentage in the different types<br />

of roads, to evidence the importance of the selected l<strong>and</strong>scape variables to make the area more<br />

or less susceptible to these types of accidents; collisions’ percentage at different time of the day<br />

<strong>and</strong> months of year (date), to underst<strong>and</strong> the relation of road type to ungulate movements <strong>and</strong><br />

behaviour.<br />

3. Results<br />

According to our data, the majority of the ungulate vehicle-collisions with tractable information<br />

occurred with wild boar (86%), followed by roe deer (9%) <strong>and</strong> lastly the red deer with 5% of the<br />

cases (Figure 1).<br />

Figure 1: Percentage of ungulate vehicle-collisions per species in central <strong>and</strong> northern Portugal.<br />

Taking into account the road type (Figure 2) it is evident that the majority of ungulate vehiclecollisions<br />

occurred at national roads (EN), with 131 cases among 186, which corresponds to<br />

70,4% of the cases, followed by municipal roads (EM) (12,9%), high speed roads (A) (9,1%),<br />

main road (IP) (5,4%) <strong>and</strong> complementary roads (IC) <strong>and</strong> forestry way (CF) with 2 accidents<br />

each one (1,1%).<br />

Figure 2: Ungulate vehicle-collisions in different road types (EN; EM; A; IP; IC; CF) in central <strong>and</strong><br />

northern Portugal<br />

The distribution of ungulate vehicle-collisions throughout the months of the year showed that<br />

the majority of accidents occurred in winter, particularly in the months of November (19,5%),<br />

December (13,3%) <strong>and</strong> January (13,8%) <strong>and</strong> in summer its visible a peak in August (10,8)<br />

(Figure 3).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Sara Marques et al. 2010. Impact of roads on ungulate species: a preliminary approach in Portugal<br />

254<br />

Figure 3: Distribution of ungulate vehicle-collisions over the months of the year (1999 to 2009) in central<br />

<strong>and</strong> northern Portugal.<br />

Day time analysis highlights the nigh period that concentrate the biggest ungulate<br />

vehicle-collisions number with 129 cases (70,9%). The accidents involving ungulates<br />

occurred in the evening showed a smaller but still relevant peak with 27 cases of the<br />

182 (14,8%) (Figure 4).<br />

The number of accidents during the dawn <strong>and</strong> day are the less expressive, corresponding<br />

both to 14,3% of the total ungulate vehicle-collisions in central <strong>and</strong> northern Portugal.<br />

Figure 4: Ungulate vehicle-collisions at different time of day (dawn; day; evening; night) in central <strong>and</strong><br />

northern Portugal (1999-2009 years interval).<br />

4. Discussion<br />

Linear infrastructures like roads <strong>and</strong> the traffic associated have impacts on wildlife <strong>and</strong><br />

particularly in this case, in ungulate species. To underst<strong>and</strong> the size of the problem it’s<br />

imperative to monitor roads kills, on a national scale, <strong>and</strong> comprehend the various parameters<br />

involved in the accidents.<br />

Our data support that in Portugal the wild boar was the most affected wild ungulate in central<br />

<strong>and</strong> northern Portugal with 86% of the vehicle-collisions, <strong>and</strong> in all districts of the study area<br />

this ungulate species was the most affected target because it’s the ungulate more widely<br />

distributed in Continental Portugal, occupying almost the entire national territory except for<br />

large urban settlements <strong>and</strong> some coastal areas (Vingada et al. 2010).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Sara Marques et al. 2010. Impact of roads on ungulate species: a preliminary approach in Portugal<br />

255<br />

Roe deer’s collisions correspond to a 9% of amount cases, which support the reduced number of<br />

animals killed by cars. Population estimates show that current populations of roe deer in<br />

Portugal include between 3000 <strong>and</strong> 5000 animals (Vingada et al. 2010), with the highest<br />

densities concentrated in the northern region of the country where the roe deer collisions were<br />

higher (districts of Braga, Viana do Castelo <strong>and</strong> Bragança).<br />

The red deer vehicle-collisions correspond to a 5% of the cases which proof the reduced <strong>and</strong><br />

scattered distribution of this animal at Portugal (Vingada et al. 2010). The data of red deer<br />

collisions were mostly in the districts of Coimbra, Viseu, Bragança, Viana do Castelo <strong>and</strong><br />

Castelo Branco which is according to dispersed Portuguese northern <strong>and</strong> central populations<br />

(Tejo Internationa, Lousã <strong>and</strong> Montesinho).<br />

According to the analysis of road type the majority of ungulate vehicle-collisions occurred at<br />

National Roads (EN), with overwhelming percentage of 70,4%, probably due to the high traffic<br />

volume <strong>and</strong> bad road conditions. The lower number of accidents with ungulate in main roads<br />

(IP), complementary roads (IC) <strong>and</strong> high speed road (A) could be explained by the existence of<br />

lateral road protections (fences) in both sides of the road which avoid the invasion of wild<br />

animals. Only two cases of wildlife vehicle-collisions were recorded in forestry ways (CF)<br />

because this type of road has not regularly traffic, although of the bad conditions.<br />

The majority of ungulate vehicle-collisions occurred in winter, particularly in November,<br />

December <strong>and</strong> January. This can be explained by the fact that the big game hunting period is<br />

mainly from October to February <strong>and</strong> it’s precisely in this time of the year that the animals are<br />

more restless (Fonseca <strong>and</strong> Correia 2008). The August ungulate-vehicle collisions’ peak can be<br />

derived from the holiday’s traffic growth in Portugal.<br />

As expected the overwhelming percentage of accidents occurred at night (70,9%) because it’s<br />

precisely the time of the day that the ungulates are more active (Apollonio et al. 2010). During<br />

the dawn, day <strong>and</strong> evening the number of accidents involving ungulates are smaller.<br />

This study indicates that there are some variables that can contribute to the ungulate vehiclecollisions<br />

increase such as the road type, time of the day <strong>and</strong> month of the year. The first<br />

variable can be amended by man in the sense of prevent <strong>and</strong> mitigate the accidents with wildlife,<br />

particularly involving large animals, such as ungulates, but the second concerns to the biology<br />

<strong>and</strong> behaviour of animals.<br />

Gradually there is a growing concern in the road planning <strong>and</strong> implementation, according to the<br />

measures to prevent <strong>and</strong> mitigate the impacts on wildlife. The design of infrastructures in the<br />

l<strong>and</strong>scape has an enormous relevance to wildlife because it causes mainly the fragmentation of<br />

available ungulate habitat <strong>and</strong> resulting in a diminished habitat connectivity <strong>and</strong> permeability.<br />

Mitigations measures that include wildlife crossing structures not only substantially reduce road<br />

mortality, but also allow movements of animals across the road (Huijser et al. 2009). Properly<br />

designed wildlife crossing structures may help to reduce the barrier effect of roads while<br />

decrease ungulate vehicle-collisions (Gagnon et al. 2007). This connectivity is essential to<br />

survival probability of the fragmented populations of some species, in determinate regions. As<br />

referred by Seiler (2004) if these measures will be combined with traffic adjustments, such as<br />

reduced speed limits, rerouting of traffic flow or improvement of features roads, traffic safety<br />

may further be improved <strong>and</strong> the collisions with wildlife will reduce.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Sara Marques et al. 2010. Impact of roads on ungulate species: a preliminary approach in Portugal<br />

256<br />

References<br />

Apollonio M., Andersen, R. & Putman, R.J. (Eds.) 2010. European Ungulates <strong>and</strong> their<br />

Management in the 21 st Century. Cambridge University Press. 618 pp.<br />

Bruinderink, G. <strong>and</strong> Hazebroek, E., 1996. Ungulate traffic collisions in Europe. Conservation<br />

Biology, 10(4): 1059-1067.<br />

Christie, J.S., <strong>and</strong> Nason, S., 2003. Analysis of ungulate-vehicle collisions on arterial highways<br />

in New Brunswick. The 31st Annual Conference of the Canadian Society for Civil<br />

Engineering, Moncton, New Brunswick, 11p.<br />

Fonseca, C. <strong>and</strong> Correira, F., 2008. O Javali: património natural transmontano. João Azevedo<br />

Editor (1ª edição), Mir<strong>and</strong>ela, Portugal, 168p.<br />

Gagnon, J., Schweinsburg, R. <strong>and</strong> Dodd, N., 2007. Effects of roadway traffic on wild ungulates:<br />

a review of the literature <strong>and</strong> case study of Elk in Arizona. In Proccedings of the 2007<br />

International Conference on Ecology <strong>and</strong> Transportation, edited by C. Leroy Irwin,<br />

Debra Nelson <strong>and</strong> K.P. Mcdermott. Raleigh, NC: Center for Transportation <strong>and</strong> the<br />

Environment, North Carolina State University, 449-458p.<br />

Huijser, M., Duffield, J., Clevenger, A., Ament, R. <strong>and</strong> McGowen, P., 2009. Cost-benefit<br />

analysis of mitigation measures aimed at reducing collisions with large ungulates in the<br />

United Satates <strong>and</strong> Canada: a decision support tool. Ecology <strong>and</strong> Society, 14(2): 15.<br />

Seiler, A., 2004. Trends <strong>and</strong> spatial patterns in ungulate-vehicle collisions in Sweden. Wildlife<br />

Biology, 10(4): 301-313.<br />

Vingada, J., Fonseca, C., Cancela, J., Ferreira, J. <strong>and</strong> Eira, C., 2010. Ungulates <strong>and</strong> their<br />

Management in Portugal. In: Apollonio, M., Andersen, R. & Putman, R.J. (Eds.)<br />

European Ungulates <strong>and</strong> their Management in the 21st Century. Cambridge University<br />

Press: 392-418.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.G.C. Martins & J. Aranha 2010. Electrical network hazard assessment for the avifauna in Portugal<br />

257<br />

Electrical network hazard assessment for the avifauna in Portugal<br />

Miguel G. C. Martins 1 & José Aranha 1,2*<br />

1 Departamento de Ciências Florestais e Arquitectura Paisagista, Universidade de Trásos-Montes<br />

e Alto Douro, 5001-801 Vila Real<br />

2 CITAB, Universidade de Trás-os-Montes e Alto Douro, 5001-801 Vila Real<br />

Abstract<br />

The suspended electrical network represents a danger to the birds’ wildlife conservation, both<br />

because of the possibility of collision <strong>and</strong> electrocution. L<strong>and</strong> use type is strictly related to<br />

birds’ presence, eating habits <strong>and</strong> nesting. This way they are factors that are directly related to<br />

collision <strong>and</strong> electrocution hazards.<br />

This work was based on the occurrences geographical position <strong>and</strong> the location of suspended<br />

electrical network. The available data were used to create a GIS in order to calculate hazard<br />

maps, by means of geostatistical processes <strong>and</strong> multivariate analysis.<br />

The final results indicate that approximately 46% of the total (km) electrical network analysed<br />

are classified with the two higher hazard classes in the case of electrocution, <strong>and</strong> about 40% in<br />

the case of collision. The classification maps show that the danger of electrocution is greatest in<br />

the central <strong>and</strong> southern Portugal, <strong>and</strong> the danger of collision is higher in the central region.<br />

Keywords: Avifauna; Collision, Electrocution; Hazard Assessment, GIS<br />

1. Introduction<br />

In the year of 2003, four Portuguese entities: the Portuguese enterprise for electrical energy<br />

(EDP – Electricidade De Portugal), the Portuguese Institute for Nature Conservation (ICN –<br />

Instituto da Conservação da Natureza), the National Association for Nature Conservation<br />

(Quercus - Associação Nacional de Conservação da Natureza) <strong>and</strong> the Portuguese Society for<br />

Birds Survey (SPEA - Sociedade Portuguesa para o Estudo das Aves), signed up a protocol, in<br />

order to analyse the relationship between the Portuguese electrical network (high <strong>and</strong> middle<br />

power) <strong>and</strong> the birds’ wildlife conservation.<br />

This research project was derived for entire continental Portugal, with special attention to areas<br />

under special protection areas (SPA’s) for birds nesting <strong>and</strong> migration (ZPE - Zonas de<br />

Protecção Especial) <strong>and</strong> areas classified Important Bird Areas (IBA’s), whish represent<br />

1372966 ha.<br />

The study sought to characterize, in general, the impacts of the electrical power network on the<br />

avifauna, in the identification <strong>and</strong> classification of power lines <strong>and</strong> their wire supports, in order<br />

to create a hazard index (Br<strong>and</strong>ão et al. 2005; Martins 2009).<br />

The results from a research project derived in Spain showed that the use of orange PVC spirals<br />

along electrical power wire network leaded to a decrease in 60% of birds’ collision to electrical<br />

wire. The results also showed that the use of these orange PVC spirals reduced the number of<br />

birds flying between suspended electrical wire leading birds to fly over or under electrical<br />

power wire network (Ferrer & Janss, 1999).<br />

* Corresponding author; Tel.: + 00 351 259 350 856 - Fax: + 00 351 250 350 480<br />

Email address : j.aranha.utad@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.G.C. Martins & J. Aranha 2010. Electrical network hazard assessment for the avifauna in Portugal<br />

258<br />

Results from a 25 year research in USA, showed that the use of non conductive materials for<br />

isolate the ground wire from the phase conductor wire or the distance enlargement between<br />

these two electrical wire, leaded to a significant decrease in birds of prey dead, such as eagles<br />

<strong>and</strong> hawk, by electrocution. Another important result is related to birds’ propensity to nest on<br />

the suspension towers of electric network. This way, the researchers proposed the use of l<strong>and</strong>ing<br />

<strong>and</strong>/or nesting platforms erected over towers ends (James & Haak 1979).<br />

The aim of this paper is to present the results from a Geographical Information System created<br />

to analyse the relationship between the Portuguese electrical network (high <strong>and</strong> middle power)<br />

<strong>and</strong> the birds’ wildlife conservation (Scott et al. 1972; Meyer 1978; James & Haak 1979;<br />

Beaulaurier 1981; Burrough 1987; Janns & Ferrer 1998; Bernhardsen 1999; Janss 2000).<br />

The study was conducted for entire continental Portugal <strong>and</strong> was developed in protected areas,<br />

IBAs <strong>and</strong> SPAs (see Figure 1), in a total area of, approximately, 1 409 365ha (Br<strong>and</strong>ão et al.<br />

2005). They are, in this territory, the most important sites for wild birds, bringing together more<br />

than 90% of the national population of at least 21 species of Annex I of Birds Directive. In order<br />

to improve the management of fieldwork, the study area was divided into 4 sampling zones <strong>and</strong><br />

data were collected along 61 senses performed on sections selected for the study of hazard rates,<br />

the frequency estimates for birds <strong>and</strong> species diversity (Br<strong>and</strong>ão et al. 2005).<br />

Figure 1: Study area location [Adapted from Portuguese Environment Atlas 2009]<br />

2. Methodology<br />

For a year, all incidents relating to birds killed or injured by the electric network, within<br />

protected areas, natural parks <strong>and</strong> national parks, have been reported, recorded its position with<br />

a GPS <strong>and</strong> recorded the nature of death or injury. The information collected was recorded in a<br />

database, which was used to create a geographic information system.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.G.C. Martins & J. Aranha 2010. Electrical network hazard assessment for the avifauna in Portugal<br />

259<br />

In a second stage, the GIS was updated with information concerning:<br />

• Portuguese electrical network<br />

• Portuguese Roads network<br />

• Settlements Location<br />

• L<strong>and</strong> use <strong>and</strong> l<strong>and</strong> cover map<br />

• Digital Elevation Model<br />

In the third stage, using 3D Analyst <strong>and</strong> Spatial Analyst Tools, the recorded information was<br />

processed in order to calculate:<br />

• L<strong>and</strong> slope <strong>and</strong> aspect<br />

• Distances to water bodies<br />

• Distance to road network<br />

The fourd stage was dedicated to geodata analysis <strong>and</strong> multivariate statistics calculation. They<br />

were used Principal Components Analysis, Cluster Analisys <strong>and</strong> Geostatistical Analysis, in<br />

order to stablish a relationship between the Portuguese electrical network (high <strong>and</strong> middle<br />

power), environmental characteristics <strong>and</strong> the birds’ wildlife conservation (Morrison, 1991;<br />

Reis, 1997; Hoef et al, 2001; Soares 2006).<br />

Results from previous present stages were, then, used to create collition <strong>and</strong> electrocution<br />

hazard maps for IBAs <strong>and</strong> SPAs<br />

3. Result<br />

Results from data management shown that the frequency of deaths is higher in Steppe l<strong>and</strong> <strong>and</strong><br />

Agroforestry areas. Tables 1 <strong>and</strong> 2 summarizes all deaths by type of habitat <strong>and</strong> electrical power<br />

line hazard level.<br />

Table 1: Deaths by type of habitat<br />

Habitat (class code) Total (n) %<br />

Steppe (1) 110 28.06<br />

Shrub l<strong>and</strong> (2) 32 8.16<br />

<strong>Forest</strong>ed l<strong>and</strong> (3) 51 13.01<br />

Inl<strong>and</strong> wetl<strong>and</strong>s (4) 16 4.08<br />

Coastal wetl<strong>and</strong>s (5) 5 1.28<br />

Agroforestry mosaic (6) 174 44.39<br />

Other (e.g. bare soil) (7) 4 1.02<br />

Total 392 100.00<br />

Results from Principal Components Analysis revealed a positive correlation between the<br />

number of deaths <strong>and</strong> the number of power line network planes <strong>and</strong> a negative correlation to<br />

distance to roads. This could indicate that the "mortality rate" increases as "distance to roads,"<br />

decreases because power line network is often parallel to road network <strong>and</strong> birds need to cross<br />

roads for feeding in around l<strong>and</strong>.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.G.C. Martins & J. Aranha 2010. Electrical network hazard assessment for the avifauna in Portugal<br />

260<br />

Table 2: Electrical power line classification, according to hazard level<br />

Electrocution<br />

Collision<br />

Class Death_class Number of lines Length (km) Length (%)<br />

1 [0 ; 0,83] 6447 1597,375 9,21<br />

2 [0,83 ; 1,66] 4386 1698,577 9,79<br />

3 [1,66 ; 2,49] 14926 6024,488 34,72<br />

4 [2,49 ; 3,33] 11267 4899,044 28,24<br />

5 [3,33 ; 4,16] 8223 3130,009 18,04<br />

1 [0 ; 1,32] 6248 2649,333 17,28<br />

2 [1,32 ; 2,65] 9360 3439,777 22,46<br />

3 [2,65 ; 3,97] 10557 3761,530 24,53<br />

4 [3,97 ; 5,30] 6365 2356,527 15,37<br />

5 [5,30 ; 6,63] 7685 3128,657 20,40<br />

Final results enable to state that Portuguese country areas with lower degree of birds’ electric<br />

shock danger are: the north <strong>and</strong> part of the center, as well as the southwestern district of Lisbon,<br />

<strong>and</strong> the coastline of the districts of Setúbal <strong>and</strong> Beja.<br />

In the districts of Guarda, Castelo Branco; Bragança, Braga, Porto, Portalegre, Santarém (North),<br />

Setubal <strong>and</strong> Viana do Castelo dominates the middle class of danger (see Figure 1 for location).<br />

The regions under high hazard level are: the districts of Évora, Santarém (South), Lisboa<br />

(northeast) Beja (central <strong>and</strong> northern) <strong>and</strong> Faro.<br />

According to the power line network characteristics <strong>and</strong> location, the south of Portugal is the<br />

results most concern as well the central coast.<br />

4. Discussion<br />

The objectives of this study were successfully achieved. The entire power line net work was<br />

classified <strong>and</strong> analysed, for the entire country, the relationship between power network <strong>and</strong><br />

environmental characteristics, with special relevance for Natural Regions <strong>and</strong> Protected Areas.<br />

We would like to highlight the following results:<br />

the worst case scenario, related to collision hazard, indicates that the high <strong>and</strong> very high hazard<br />

classes of collision are present in more than 36% of the total analysed kilometres,<br />

for electrocution hazard, the worst forecast indicates the prevalence of high <strong>and</strong> very high<br />

hazard classes in approximately 50% of the total analysed kilometres.<br />

These results are worrying in terms of ecology, <strong>and</strong> should wherever possible, be eliminated or<br />

minimised in order to reduce negative effects on avifauna,<br />

for Natural Regions <strong>and</strong> Protected Areas, the results indicate that the high <strong>and</strong> very high hazard<br />

classes of collision are present in more than 43% of the total analysed kilometres, the high <strong>and</strong><br />

very high hazard classes of electrocution are present in more than 34% of the total analysed<br />

kilometres <strong>and</strong> 43% classified as middle class.<br />

References<br />

Bernhardsen, T. 1999. Geographical Information Systems, an introduction – Second<br />

Edition. John Wiley & Sons, Inc. New York. U.S.A.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.G.C. Martins & J. Aranha 2010. Electrical network hazard assessment for the avifauna in Portugal<br />

261<br />

Beaulaurier, D. L. 1981. Mitigation or bird collision with transmission lines. Bonneville<br />

Power Administration. U.S. Department of Energy. Portl<strong>and</strong>. U.S.A<br />

Burrough, P. A. 1987. Principles of geographical information systems for l<strong>and</strong><br />

resources assessment. Oxford, Clarendon Press. Oxford. United Kingdom.<br />

Br<strong>and</strong>ão, R., Infante, S., Ministro, J., Neves, L. 2005. Estudo sobre o Impacto das<br />

Linhas Eléctricas de Média e Alta Tensão na Avifauna em Portugal. Quercus<br />

Associação Nacional de Conservação da Natureza e SPEA Sociedade Portuguesa<br />

para o Estudo das Aves. Castelo Branco. Portugal.<br />

Ferrer, M., Janss, G. F. E. 1999. Aves y Líneas Eléctricas, Colisión, Electrocución y<br />

Nidificación. Serviços Informativos Ambientales - Quercus. Madrid. Espanha.<br />

Hoef, J. M., Johnston, K., Lucas, N., Krivoruchko, K. 2001. Using ArcGIS<br />

Geostatistical Analyst. ESRI. USA.<br />

Janss, G. F. 2000. Avian mortality from power lines: a morphologic approach of a<br />

species-specific mortality. Biological Conservation 95: 353-359.<br />

James, B. W. & Haak, B.A. 1979. Factors affecting avian flight behavior <strong>and</strong> collision<br />

mortality at transmission lines. Bonnevile Power Administration Report. U.S.<br />

Department of Energy. Oregon. U.S.A.<br />

Janns, G. F. & Ferrer, N. 1998. Rate of collision with power lines: conductor-marking<br />

<strong>and</strong> groundwire-marking. Journal of Field Ornithology 69: 8-17.<br />

Meyer, J. R. 1978. Effects of transmission lines on bird flight behavior <strong>and</strong> collision<br />

mortality. Bonnevile Power Administration Report. U. S. Department of Energy.<br />

Oregon. U.S.A.<br />

Soares, A. 2006. Geoestatística para as ciências da terra e do ambiente – Segunda<br />

Edição. Instituto superior Técnico. Lisboa. Portugal.<br />

Acknowledgement<br />

Authors woul like to expresse is acknowledge to Fundação para a Ciência e Tecnologia (FCT),<br />

project REEQ/1163/AGR/2005, CITAB – UTAD, EDP – Electricidade De Portugal, the<br />

Portuguese Institute for Nature Conservation (ICN – Instituto da Conservação da Natureza), the<br />

National Association for Nature Conservation (Quercus - Associação Nacional de Conservação<br />

da Natureza) <strong>and</strong> the Portuguese Society for Birds Survey (SPEA - Sociedade Portuguesa para o<br />

Estudo das Aves).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Modeling l<strong>and</strong> use/cover change <strong>and</strong> biodiversity conservation in Mexico<br />

262<br />

Modeling l<strong>and</strong> use/cover change <strong>and</strong> biodiversity conservation in<br />

Mexico<br />

Jean-François Mas 1* , Azucena Pérez Vega 2 , Keith Clarke 3 & Víctor Sánchez-Cordero 4<br />

1 Centro de Investigaciones en Geografía Ambiental - Universidad Nacional Autónoma<br />

de México, Mexico<br />

2 Departamento de Ingeniería Civil - Universidad de Guanajuato, Mexico<br />

3 Department of Geography – University of California - Santa Barbara, USA<br />

4 Instituto de Biología - Universidad Nacional Autónoma de México, Mexico<br />

Abstract<br />

A nationwide multidate GIS database was generated in order to carry out the quantification <strong>and</strong><br />

spatial characterization of l<strong>and</strong> use/cover changes (LUCC) in Mexico during the last decades.<br />

Digital maps from three different dates (1993, 2002 <strong>and</strong> 2007) were revised <strong>and</strong> integrated into<br />

a GIS database along with ancillary data (Road network, settlements, slope <strong>and</strong> socioeconomical<br />

parameters at the municipality level). L<strong>and</strong> cover maps were overlaid in order to<br />

generate LUCC maps <strong>and</strong> calculate two indices de LUCC: 1) a simple rate of deforestation <strong>and</strong><br />

2) a rate of degradation which takes into account the degradation <strong>and</strong> recovery processes. An<br />

analysis of causes <strong>and</strong> drivers of LUCC was conducted, at the municipality level, computing the<br />

Spearman coefficient between these two rates <strong>and</strong> biophysical <strong>and</strong> socio-economic factors.<br />

<strong>Change</strong> trends were also compared with biodiversity distribution. Preliminary results show that<br />

although rates of deforestation have decreased during the most recent period, LUCC still<br />

represents a serious threat to biodiversity conservation in Mexico.<br />

Keywords:l<strong>and</strong> use/cover change, modeling, drivers, biodiversity<br />

1. Introduction<br />

Mexico is a megadiverse country, but biodiversity is threatened by the loss of native vegetation<br />

due to the high rates of deforestation (FAO, 2001). Various studies have attempted to assess<br />

l<strong>and</strong> use/ cover change (LUCC) over the last decades (Mas et al., 2004). Fuller et al. (2007)<br />

examined the effect of LUCC on the distributions of 86 endemic mammal species in 1970,<br />

1976, 1993, <strong>and</strong> 2000 in Mexico. They showed that this fauna could have been protected<br />

considerably more economically if a conservation plan had been implemented in 1970 than is<br />

possible today due to extensive conversion of primary habitats. At each time step, optimal<br />

conservation area networks were selected to represent all species. These authors found that 90%<br />

more l<strong>and</strong> must be protected after 2000 to protect adequate mammal habitat than would have<br />

been required in 1970. The goals of this study are to 1) delineate maps of LUCC in Mexico for<br />

different periods, 2) identify variables <strong>and</strong> drivers that influence the LUCC rates <strong>and</strong> 3) compare<br />

LUCC trends with a biodiversity map to evaluate the possible effects of LUCC on biodiversity<br />

conservation. In this paper, we present the preliminary results.<br />

2. Methodology<br />

* Corresponding author. Tel.: +52 443 328 38 35 - Fax: +52 443 38 80<br />

Email address: jfmas@ciga.unam.mx<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Modeling l<strong>and</strong> use/cover change <strong>and</strong> biodiversity conservation in Mexico<br />

263<br />

2.1. Material<br />

The following data were used:<br />

• Maps of l<strong>and</strong> use/cover (LUC) at 1:250,000 scale from the National Institute of<br />

Geography, Statistics <strong>and</strong> Informatics (INEGI) for 1993, 2002 <strong>and</strong> 2007. These maps are<br />

compatible with regards to scale <strong>and</strong> classification scheme. The classification scheme<br />

distinguishes primary covers <strong>and</strong> 3 categories of secondary l<strong>and</strong> covers (with herbaceous,<br />

scrub <strong>and</strong> tree secondary vegetation respectively). According to INEGI, primary<br />

vegetation is defined as relatively undisturbed vegetation that preserves, in large part, its<br />

condition of density, coverage, <strong>and</strong> species composition from its original, primary,<br />

ecosystem. Secondary vegetation is defined as the vegetation which substitutes totally or<br />

partially the original (primary) vegetation as a result of secondary succession.<br />

• Map of species richness: To generate this map, Sánchez-Cordero et al. (2005) modeled<br />

ecological niches for the 459 continental mammal species of Mexico using point<br />

occurrence distribution from national <strong>and</strong> international scientific collections <strong>and</strong><br />

environmental data layers, including potential vegetation type, elevation, topography <strong>and</strong><br />

climatic parameters using the Genetic Algorithm for Rule-set Prediction (Stockwell et al.<br />

1999; Anderson et al. 2003). Then maps of each species were overlain in order to<br />

calculate the species richness.<br />

• Maps of ancillary data (digital elevation model, roads maps, human settlements, municipal<br />

boundaries).<br />

• Socio-economic data from the INEGI organized by municipality (Population census for<br />

2000 <strong>and</strong> 2005).<br />

GIS operations were carried out with the program ArcGIS (ESRI, Redl<strong>and</strong>s, CA) <strong>and</strong> statistical<br />

analysis <strong>and</strong> graphs were created using R (R Development Core Team 2009).<br />

2.2. LUCC Monitoring<br />

Mapping of LUCC was done by overlaying the LUC maps of different dates. Based upon<br />

LUCC maps, areas of change were tabulated <strong>and</strong> rates of change, including the rate of<br />

deforestation, were computed. As the rate of deforestation is sensitive to the change from forest<br />

areas (primary <strong>and</strong> secondary covers altogether) to non forest area only, we also applied an<br />

“index of conservation” which takes into account forest degradation <strong>and</strong> recovery (e.g.<br />

transitions between secondary categories). L<strong>and</strong> cover categories were associated to a weight<br />

value ranging from 0 to 4 for anthropogenic, herbaceous secondary, scrub secondary, tree<br />

secondary <strong>and</strong> primary forest forest respectively. Mean values of this index was calculated for<br />

2004 <strong>and</strong> 2007 for each municipality.<br />

3.2. Relationship between LUCC <strong>and</strong> socioeconomic features<br />

To determine which socioeconomic factors are most likely to be indirect drivers of deforestation<br />

we calculated the rate of deforestation <strong>and</strong> the variation of the conservation index for each<br />

municipality for 2004-2007 <strong>and</strong> compared them with various indices describing population<br />

density, education, poverty <strong>and</strong> accessibility to resources. These indices were: a) Population<br />

density in 2000 <strong>and</strong> 2005 (people per km 2 ) <strong>and</strong> the variation of density between these two dates;<br />

b) settlements density (number of settlements per km 2 ); c) proportion of the population older<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Modeling l<strong>and</strong> use/cover change <strong>and</strong> biodiversity conservation in Mexico<br />

264<br />

than 12 years without primary education; d) proportion of the population speaking an<br />

indigenous language; e) proportion of the population living in small settlements (with less than<br />

100 <strong>and</strong> 2500 inhabitants); f) proportion of population between 20 <strong>and</strong> 39 years (%), which is<br />

expected to be inversely correlated with migration; g) proportion of houses with a cement roof<br />

as an index of social welfare; h) the Gini index which measures inequality <strong>and</strong> ranges<br />

theoretically from 0 to 100, where 0 is perfect equality <strong>and</strong> 100 perfect inequality; i) the mean<br />

salary (expressed as the number of minimum wage salaries) <strong>and</strong> the proportion of the<br />

population with less than one <strong>and</strong> two minimum wage salaries; j) the natural cover area (ha) <strong>and</strong><br />

the proportion of total area covered by natural cover; k) the mean slope (degrees); <strong>and</strong> l) the<br />

road density (km of road per km 2 ).<br />

3. Result<br />

3.1. LUCC Monitoring<br />

Figure 1 <strong>and</strong> table 1 shows a significant decrease of forest area (except secondary temperate<br />

forest) <strong>and</strong> an increase of crop <strong>and</strong> pasture l<strong>and</strong>s during both periods. However, rates of change<br />

are lower during the more recent period.<br />

Figure 1: Areas of the main l<strong>and</strong> cover types in 1993, 2002 <strong>and</strong> 2007 (km 2 )<br />

Table 1: Rates of changes for the main l<strong>and</strong> cover types<br />

L<strong>and</strong> cover category Area (km 2 ) <strong>Change</strong> (km 2 /yr) Rate of change<br />

(%/yr)<br />

1993 2002 2007 1993-2002 2002-07 1993-2002 2002-07<br />

Arid tropical scrub 571383 558297 554661 -1454 -727 -0.26 -0.13<br />

Crop l<strong>and</strong>s 278424 308592 321597 3352 2601 1.15 0.83<br />

Pasture l<strong>and</strong>s 172278 187587 188964 1701 275 0.95 0.15<br />

Primary Temperate <strong>Forest</strong> 270432 252891 247707 -1949 -1037 -0.74 -0.41<br />

Primary Tropical <strong>Forest</strong> 233388 229815 227205 -397 -522 -0.17 -0.23<br />

Secondary Temperate <strong>Forest</strong> 78021 89028 93987 1223 992 1.48 1.09<br />

Secondary Tropical <strong>Forest</strong> 116316 99333 93726 -1887 -1121 -1.74 -1.16<br />

3.2. Analysis of drivers<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Modeling l<strong>and</strong> use/cover change <strong>and</strong> biodiversity conservation in Mexico<br />

265<br />

For the statistical analysis we use only the municipalities with a scrubl<strong>and</strong>s or forest area<br />

covering at least 500 ha <strong>and</strong> 30% of the municipality. 2314 municipalities (of a total of 2443)<br />

fulfilled this condition <strong>and</strong> represent more than 96% of the forest <strong>and</strong> scrub area of the country.<br />

Figure 2 shows the rate of deforestation per municipality.<br />

Table 2 shows that rate of deforestation <strong>and</strong> the variation of the conservation index are strongly<br />

associated (R = 0.77, p < 0.01) <strong>and</strong> that both indices are weakly, but significantly, related to<br />

some of the indices describing the socio-economic <strong>and</strong> environmental characteristics of the<br />

municipalities. Unexpectedly, population density is negatively correlated with degradation <strong>and</strong><br />

deforestation during 2002-2007 although the increase of density is related with an increase of<br />

deforestation. Indices related with poverty present a positive correlation with deforestation <strong>and</strong><br />

degradation. No significant correlation was found with the Gini index. Higher slopes <strong>and</strong><br />

unexpectedly road density tend to reduce the deforestation/degradation.<br />

Index<br />

Table 2: Spearman correlation between change <strong>and</strong> municipality characteristics<br />

Rate of<br />

deforestation*<br />

Degradation (Variation of<br />

conservation index) *<br />

R p R p<br />

Population density 2000 (people per km 2 ) -0.04 0.14 -0.05 0.03<br />

Population density 2005 (people per km 2 ) -0.02 0.34 -0.05 0.06<br />

Population density variation 2000-2005 0.11 0.00 0.05 0.05<br />

Settlements density (settlements per km 2 ) -0.07 0.01 -0.06 0.02<br />

Population older than 12 years without primary -0.10 0.00 -0.05 0.06<br />

education (%)<br />

Population speaking an indigenous language (%) -0.04 0.07 -0.06 0.02<br />

Population living in settlement of less than 100 -0.02 0.37 0.00 0.97<br />

inhabitants (%)<br />

Population living in settlement of less than 2500 -0.09 0.00 -0.05 0.03<br />

inhabitants (%)<br />

Population between 20 <strong>and</strong> 39 years (%) 0.17 0.00 0.06 0.02<br />

Houses with cement roof (%) 0.06 0.01 0.03 0.21<br />

Gini index 0.00 0.95 0.00 1.00<br />

Mean salary (number of minimum salary) 0.13 0.00 0.13 0.00<br />

Population with less than one minimum salary (%) -0.11 0.00 -0.12 0.00<br />

Population with less than 2 minimum salary (%) -0.13 0.00 -0.13 0.00<br />

Natural cover area (ha) 0.23 0.00 0.23 0.00<br />

Natural cover area (%) 0.09 0.00 0.11 0.00<br />

Mean slope (degrees) -0.24 0.00 -0.15 0.00<br />

Road density (km/km 2 ) -0.13 0.00 -0.11 0.00<br />

Rate of deforestation - - 0.77 0.00<br />

* A negative value for the rate of deforestation <strong>and</strong> degradation indicates recovery<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Modeling l<strong>and</strong> use/cover change <strong>and</strong> biodiversity conservation in Mexico<br />

266<br />

Figure 2: Rates of deforestation by municipality (2002-2007)<br />

Spearman’s coefficient of rank correlation between the rate of deforestation <strong>and</strong> the average<br />

species richness per municipality is 0.06 (p < 0.01). The coefficient between the rate of<br />

degradation <strong>and</strong> the average species richness is 0.11 (p < 0.01) which indicates that most<br />

threatened areas tend to present more biodiversity.<br />

4. Discussion<br />

According to INEGI maps, the patterns of change observed during 1993-2002 remain similar in<br />

2002-2007: Crop <strong>and</strong> pasture area is increasing <strong>and</strong> forest area, expect secondary temperate<br />

forest, is decreasing. However, the rates of deforestation are lower during the second period<br />

except for primary tropical forest, which presented an increase of the rate of deforestation. Rates<br />

of deforestation <strong>and</strong> degradation tend to be higher in biodiverse areas <strong>and</strong>, therefore, LUCC still<br />

represents a serious threat to biodiversity conservation in Mexico.<br />

The results of the analysis of the drivers must be taken with caution for various reasons:<br />

1) Analysis was based upon average characteristics of the municipality, yet their size is very<br />

variable (125 to more than 5,000,000 ha) <strong>and</strong> they are often heterogeneous; 2) the causes of<br />

LUCC in a given municipality are not necessarily reflected in the characteristics of this<br />

municipality (spatial lag); 3) LUCC are likely due to different processes over the entire territory,<br />

which can obscure meaningful explanatory variables in a nationwide study; 4) only recent<br />

LUCC are observed, in many settled regions, with high population <strong>and</strong> road density, rates of<br />

deforestation are low because few forests remain. Moreover, due to multicollinearity, a variable<br />

correlated with rates of LUCC may actually have no influence <strong>and</strong> be correlated with the true<br />

causal variables. In further research, method such as hierarchical partitioning will be used in<br />

order to deal with these limitations (MacNally 2000; 2002).<br />

However, the results obtained which indicate that marginal poor areas present lower rates of<br />

deforestation <strong>and</strong> degradation are not surprising. Many previous studies reported that most<br />

conserved natural areas in Mexico are often located in poor rural areas <strong>and</strong>/or community l<strong>and</strong>s<br />

where people have demonstrated for centuries that they have the ingenuity to cope with major<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Modeling l<strong>and</strong> use/cover change <strong>and</strong> biodiversity conservation in Mexico<br />

267<br />

environmental <strong>and</strong> social threats to their live-hoods (Klooster 2000; Alix-Garcia et al. 2005;<br />

Merino 2007, Figueroa et al. 2009, García-Barrios et al. 2009).<br />

5. Acknowledgements<br />

This research has been supported by the Consejo Nacional de Ciencia y Tecnología -<br />

CONACyT (Sabbatical <strong>and</strong> posdoctoral stays at the University of California - Santa Barbara of<br />

the first <strong>and</strong> second author respectively) <strong>and</strong> the Dirección General de Asuntos del Personal<br />

Académico (DGAPA) at the Universidad National Autónoma de México.<br />

References<br />

Alix-Garcia, J., de Janvry, A. <strong>and</strong> Sadoulet, E., 2005. A Tale of Two Communities: Explaining<br />

Deforestation in Mexico. World Development, 33(2): 219-235.<br />

Anderson, R. P., Lew, D, Peterson, A. T., 2003. Evaluating predictive models of species<br />

distributions: criteria for selecting optimal models. Ecological Modelling (162) 211-232.<br />

FAO, 2001. <strong>Global</strong> resources assessment. <strong>Forest</strong>ry paper 140.<br />

Figueroa, F., Sánchez-Cordero, V., Meave, J.A. <strong>and</strong> Trejo, I., 2009. Socioeconomic context of<br />

l<strong>and</strong> use <strong>and</strong> l<strong>and</strong> cover change in Mexican biosphere reserves. Environmental<br />

Conservation, 36 (3): 180–191.<br />

Fuller, T. M., Sánchez-Cordero, V., Illoldi-Rangel, V., Linaje, M. <strong>and</strong> Sarkar, S., 2007. The cost<br />

of postponing biodiversity conservation in Mexico. Biological Conservation, 134(4): 593-<br />

600.<br />

García-Barrios, L., Galván-Miyoshi, Y. M., Valdivieso Pérez, I. A., Masera, O. R., Bocco, G.,<br />

<strong>and</strong> V<strong>and</strong>ermeer, J., 2009. Neotropical forest conservation, agricultural intensification<br />

<strong>and</strong> rural outmigration: The Mexican Experience. BioScience, 59(10): 863-873.<br />

Klooster, D., 2000. Beyond Deforestation: The Social Context of <strong>Forest</strong> <strong>Change</strong> in Two<br />

Indigenous Communities in Highl<strong>and</strong> Mexico. Journal of Latin American Geography, 26:<br />

47-59.<br />

Mac Nally, R., 2000. Regression <strong>and</strong> model building in conservation biology, biogeography <strong>and</strong><br />

ecology: the distinction between <strong>and</strong> reconciliation of ’predictive’ <strong>and</strong> ’explanatory’<br />

models. Bio- diversity <strong>and</strong> Conservation, 9: 655–671.<br />

Mac Nally, R., 2002. Multiple regression <strong>and</strong> inference in ecology <strong>and</strong> conservation biology:<br />

further comments on identitying important predictor variables. Biodiversity <strong>and</strong><br />

Conservation, 11: 1397–1401.<br />

Mas, J.F., Velázquez, A., Díaz-Gallegos, J.R.,Mayorga-Saucedo, R., Alcántara, C., Bocco, G.<br />

Castro, R., Fernández, T., <strong>and</strong> Pérez-Vega, A., 2004. Assessing l<strong>and</strong>/use cover changes: a<br />

nationwide multidate spatial database for Mexico. International Journal of Applied Earth<br />

Observation Geoinformatics, 5: 249–261.<br />

Merino, L., 2008. Conservación comunitaria en la cuenca alta del Papaloapan, Sierra Norte de<br />

Oaxaca. Nueva Antropología, Revista de Ciencias Sociales, I68: 37-49.<br />

R Development Core Team, 2009. R: A language <strong>and</strong> environment for statistical computing. R<br />

Foundation for Statistical Computing,Vienna, Austria, URL http://www.R-project.org.<br />

Sánchez-Cordero, V., Cirelli, V., Munguia, M. <strong>and</strong> Sarkar, S., 2005. Place prioritization for<br />

biodiversity representation using species ecological niche modeling. Biodiversitic<br />

Informatics, 2: 11-23.<br />

Stockwell, D. R. B. <strong>and</strong> Peters, D., 1999. The GARP modelling system: problems <strong>and</strong> solutions<br />

to automated spatial prediction. International Journal of Geographical Information<br />

Science, 13: 143-158.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & T.K. Pereira 2010. Evaluating an old sustainable national forest in south Brazil<br />

268<br />

Evaluating an old sustainable national forest in south Brazil to decide<br />

their conservation status<br />

Rosemeri Segecin Moro 1* & Tiaro Katu Pereira 2<br />

1 DEBIO- Universidade Estadual de Ponta Grossa, PR, Brazil<br />

2 Mestrado em Gestão do Território- Universidade Estadual de Ponta Grossa, PR, Brazil<br />

Abstract<br />

National <strong>Forest</strong>s were settled in the past in order to increase forestry surveys in Brazil. After<br />

their transformation in units of conservation of sustainable use, old ab<strong>and</strong>oned sets becoming<br />

important remainders of the endangered Atlantic <strong>Forest</strong>. The recent government intention to<br />

exploit their timber has caused discussions regarding this legallity. To elucidate their appliance,<br />

reforestation in the Açunguí National <strong>Forest</strong> were evaluated. Studies showed 61 species <strong>and</strong> 30<br />

families, which the most representative were Flacourtiaceae, Lauraceae <strong>and</strong> Myrtaceae.<br />

Frequent species were Araucaria angustifolia, Cordyline dracaenoides, Cyathea corcovadensis,<br />

Casearia sylvestris, Allophyllus edulis, Clethra scabra, Dalbergia brasiliensis, <strong>and</strong> Matayba<br />

elaeagnoides. The IVI varied between 14.3 <strong>and</strong> 7.5. Diversity of species was high for secondary<br />

forested areas - Shannon’s index (H’= 3.15); evenness (J’= 0,77). There were endangered<br />

species <strong>and</strong> bioindicators that enhance the ecological importance of this National <strong>Forest</strong>. Thus,<br />

this one should not be exploited despite their original purpose.<br />

Keywords: Araucaria forest – Atlantic <strong>Forest</strong> – planted forests<br />

1. Introduction<br />

Remnants of Ombrophylous Mixed <strong>Forest</strong> is one of the most endangerous at the Atlantic<br />

<strong>Forest</strong> Biome (Castella <strong>and</strong> Britez 2004). Native species reforestation was a federal politics in<br />

Brazil in the decades of 1940-60, settling National <strong>Forest</strong>s all over the country. After the<br />

project’s dismounting <strong>and</strong> their transformation in units of conservation of sustainable use<br />

(SNUC 2000), old sets presenting vigorous sub-forests becoming important remainders of the<br />

endangered Atlantic <strong>Forest</strong>. Nowadays, the government intention to exploit their timber has<br />

caused intense internal discussions once they could be protected by law as mature regenerated<br />

forests. Intending to elucidate their appliance, we analyze sub-sets under reforestation <strong>and</strong> under<br />

natural regeneration.<br />

2. Methodology<br />

Surveys were carried out in seventeen 0.01-ha plots laid out in a 30-yr-old 400 ha planted<br />

Araucaria st<strong>and</strong> <strong>and</strong> a natural 320 ha Araucaria forest (see figure 1) at Açungui National <strong>Forest</strong><br />

in Campo Largo, State of Paraná, Southern Brazil (25 o 10'41"S - 25 o 14'18"W). The dominant<br />

soil types in the region are Inceptisol <strong>and</strong> Ultisol. The climate is classified as Cfb (under<br />

Koeppen’s climate classification system), at elevations between 640 m <strong>and</strong> 905 m above sea<br />

level. Phytosociological parameters were performed by Fitopac Software (Shepherd 1994) <strong>and</strong><br />

the samples are in the HUPG herbarium at Universidade Estadual de Ponta Grossa.<br />

* Corresponding author. Tel.:55-42-3223-4417 - Fax:55-42-3224-4747<br />

Email address: rsmoro@uepg.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & T.K. Pereira 2010. Evaluating an old sustainable national forest in south Brazil<br />

269<br />

Figure 1: Sampling sites at Açungui National <strong>Forest</strong>, Campo Largo (Paraná, Brazil).<br />

3. Results<br />

Among the l<strong>and</strong>scape units, the Natural <strong>Forest</strong> has 38% of natural Araucaria forests <strong>and</strong><br />

56% of planted Araucaria st<strong>and</strong>s. Species composition <strong>and</strong> several forest structure parameters<br />

are summarized in Tables 1 <strong>and</strong> 2. Both natural <strong>and</strong> planted Araucaria st<strong>and</strong>s showed similar<br />

phytosociological parameters. Therefore, there is evidences that ab<strong>and</strong>oned planted Araucaria<br />

st<strong>and</strong>s are able to restore this type of community in the Atlantic <strong>Forest</strong> Biome almost like the<br />

espontaneous secondary sucession.<br />

More than a half of the species found in both planted <strong>and</strong> natural Araucaria st<strong>and</strong>s were<br />

represented by Flacourtiaceae, Lauraceae <strong>and</strong> Myrtaceae. Sapindaceae <strong>and</strong> Euphorbiaceae were<br />

found only in the natural Araucaria st<strong>and</strong>. The most abundant families found under the canopy<br />

were Agavaceae, Flacourtiaceae, Sapindaceae, Cyatheaceae, Myrtaceae, Canellaceae, <strong>and</strong><br />

Euphorbiaceae. In terms of the Importance Value Index (IVI), the Flacourtiaceae, Agavaceae,<br />

Cyatheaceae, Sapindaceae, <strong>and</strong> Fabaceae were the most important families found under planted<br />

Araucaria st<strong>and</strong>s while Sapindaceae, Myrtaceae, Flacourtiaceae, <strong>and</strong> Araucariaceae were the<br />

most important families found under the natural forest canopy.<br />

In the planted Araucaria st<strong>and</strong>s only Araucaria angustifolia stood up from the canopy,<br />

otherwise in the natural forest Matayba elaeagnoides could emerge to. The more frequent<br />

species were Casearia sylvestris, Casearia lasiophylla, Casearia obliqua, Casearia<br />

inaequilatera, Vitex megapotamica, Guatteria australis, Cabralea canjerana, Cedrella fissilis,<br />

Allophyllus edulis, Cupania vernalis, Clethra scabra, Dalbergia brasiliensis, Matayba<br />

elaeagnoides, Cordyline dracaenoides, Cyathea corcovadensis, Cyathea schanschin, <strong>and</strong><br />

Myrcia rostrata.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & T.K. Pereira 2010. Evaluating an old sustainable national forest in south Brazil<br />

270<br />

Table 1: Phytosociological parameters of Araucaria forests at Açungui National <strong>Forest</strong>, Campo Largo<br />

(Paraná, Brazil).<br />

Under planted Araucaria canopy Under natural Araucaria canopy<br />

Number of tree species 61 54<br />

Number of families 30 21<br />

Total number of plants 400 160<br />

Density (plant/ha) 3.333 1.680<br />

Basal area /ha 54.867 -<br />

Canopy – mean Dbh (cm) 11.18 11.32<br />

Canopy – mean height (m) 7.12 6.82<br />

Shannon-Wienner Index 3.15 3,43<br />

Evenness 0.77 0.86<br />

Simpson Index 0.93 0.95<br />

Table 2: Species list of both planted <strong>and</strong> natural Araucaria st<strong>and</strong>s at Açungui National <strong>Forest</strong> in Campo<br />

Largo (Paraná-Brazil).<br />

Family Species Common<br />

name<br />

Ecological<br />

group*<br />

Density<br />

**<br />

Use<br />

***<br />

1 Agavaceae Cordyline dracaenoides Uvarana Pi, Si, St VC Fo<br />

Kunth<br />

2 Anacardiaceae Schinus terebinthifolius Aroeiravermelha<br />

Pi, Si, St NC Ti/<br />

Raddi<br />

Me<br />

3 Annonaceae Guatteria australis A.St. Embiú<br />

NC<br />

Hill.<br />

4 Aquifoliaceae Ilex dumosa Reiss. Caúna-miúda Pi, Si, St Fo<br />

Ilex paraguariensis A. Erveira; ervamate<br />

Pi, Si, St C Fo<br />

St.Hill.<br />

Ilex theezans Mart. Congonhagraúda<br />

Pi, Si, St<br />

Fo<br />

5 Araucariaceae Araucaria angustifolia Araucária Pi, Si, St VC Fo<br />

(Bert.) Kuntze<br />

6 Arecaceae Syagrus romanzoffiana Jerivá Pi, Si, St Fo<br />

(Cham.) Glassman<br />

7 Asteraceae Piptocarpha angustifolia Vassourãobranco<br />

Pi, Si, St<br />

Ti<br />

Dusén ex Malme<br />

8 Bignoniaceae Jacar<strong>and</strong>a micrantha Carobão Si<br />

Cham.<br />

9 Bombacaceae Chorisia speciosa St. Hill. Paineira Si, St R Ha<br />

10 Borraginaceae Cordia ecalyculata Vell. Louro-mole Si, St<br />

11 Caesalpinaceae Apuleia leiocarpa (Vogel) Grápia Si, St, Cl R<br />

J.F. Macbr.<br />

12 Canellaceae Capsicodendron dinisii Pimenteira Pi, Si, St Me<br />

(Schwanke) Occh.<br />

13 Celastraceae Maytenus evonymoides Coração-debugre<br />

Pi, Si, St VC<br />

Reiss.<br />

14 Clethraceae Clethra scabra Pers. Guaperê Pi, Si, St VC Ti<br />

15 Cunoniaceae Lamanonia ternata Vell. Guaraperê Si<br />

16 Cyatheaceae Cyathea corcovadensis Xaxim Si, St VC Ha<br />

Rad.<br />

Cyathea schanschin Mart. Xaxim Si, St C Ha<br />

17 Erythroxylaceae Erythroxylum deciduum Cocão Pi, Si R<br />

St. Hill.<br />

18 Euphorbiaceae Actinostemon concolor Laranjeira-do- Pi, Si, St R<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & T.K. Pereira 2010. Evaluating an old sustainable national forest in south Brazil<br />

271<br />

Spr.<br />

mato<br />

Sapium gl<strong>and</strong>ulatum Pau-leiteiro Pi, Si, St Me<br />

(Vell.) Pax<br />

Sebastiania brasiliensis Tajuvinha Pi, Si, St Ti<br />

Spr.<br />

Sebastiania<br />

Branquinho Pi, Si, St R<br />

commersoniana (Baill.) L.<br />

B. Sm. et Downs<br />

19 Fabaceae Dalbergia brasiliensis Farinha-seca; Pi, Si<br />

C<br />

Vog.<br />

marmeleiro<br />

Machaerium stipitatum Sapuva Pi, Si Ti<br />

Vog.<br />

20 Flacourtiaceae Banara tomentosa Clos. Guaçatungabranca<br />

Pi, Si, St R<br />

Casearia dec<strong>and</strong>ra Jacq. Guaçatungapreta;<br />

Pi, Si, St<br />

Me<br />

cambroé<br />

Casearia inaequilatera Guaçatungamiúda<br />

Pi, Si, St<br />

Camb.<br />

Casearia lasiophylla Guaçatungagraúda<br />

Pi, Si, St VC<br />

Sleumer<br />

Casearia obliqua Spr. guaçatungavermelha<br />

Pi, Si, St<br />

Me<br />

Casearia sp<br />

R<br />

Casearia sylvestris Sw. Cafezeiro Pi, Si, St VC Me<br />

Xylosma ciliatifolium Sucará Si, St NC<br />

(Clos) Eichler<br />

21 Lauraceae Cinnamomum sellowianum Canela-raposa Pi, Si, St NC<br />

(Nees) Kostern.<br />

Cryptocaria<br />

Canela-fogo St C Ti<br />

aschersoniana Mez<br />

Nect<strong>and</strong>ra lanceolata Canelaamarela<br />

Pi, Si, St C<br />

Ness et Mart. ex Ness<br />

Nect<strong>and</strong>ra megapotamica Canela-preta; Pi, Si, St<br />

(Spr.) Mez<br />

canela<br />

imbuia<br />

Ocotea dyospyrifolia Canela-goiaba Si, St<br />

(Meins.) Mez<br />

Ocotea lancifolia (Nees) Canela St, Cl<br />

Mez<br />

Ocotea puberula (A.Rich) Canela-sebo Pi, Si, St Ti<br />

Nees<br />

22 Meliaceae Cabralea canjerana (Vel.) Canjerana Si, St Ti<br />

Mart.<br />

Cedrella fissilis Vell. Cedro-rosa Pi, Si, St TiMe<br />

Trichilia elegans A. juss. Catiguá-deervilha<br />

St<br />

Trichillia catiguá A. Juss. catiguá St<br />

23 Mimosaceae Piptadenia gonoacantha Pau-jacaré Pi, Si, St NC<br />

(Mart.) Macbr.<br />

24 Monnimiaceae Mollinedia clavigera Tull. capixim Pi, Si, St C<br />

25 Moraceae Ficus guaranítica Schodat Figueira R<br />

Sorocea bonpl<strong>and</strong>i (Baill.) cincho Si, St, Cl NC<br />

W. C. Burger<br />

26 Myrsinaceae Myrsine umbellata Mart. capororocão Pi, Si, St C Ti<br />

27 Myrtaceae Campomanesia guabiroba Guabiroba Pi, Si, St C Fo<br />

(DC) Kiaersk.<br />

Eugenia blastantha<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & T.K. Pereira 2010. Evaluating an old sustainable national forest in south Brazil<br />

272<br />

Lam.<br />

32 Sapindaceae Allophylus edulis (A.St.-<br />

Hil., Cambess. & A. Juss.)<br />

Radlk).<br />

(O.Berg.) Legr.<br />

Eugenia uniflora L. pitangueira Pi, Si, St C Fo<br />

Gomidesia affinis (Camb.) guamirim<br />

Legr.<br />

Mosiera prismatica (D. Cerninho; Pi, Si, St<br />

Fo<br />

Legr.) L<strong>and</strong>rum<br />

cambuí<br />

Myrcia hatschbachii Legr. Caingá Si<br />

Myrcia myrcioides Guamirim Si<br />

(Camb.) O. Berg.<br />

Myrcia obtecta (Berg.) Guamirimbranco;<br />

Si<br />

Kiaersk.<br />

cambuí<br />

Myrcia rostrata DC Guamirimchorão<br />

Pi, Si, St<br />

Ti<br />

Myrcianthes pungens (O. Guabijú St<br />

Berg) D. Legr<strong>and</strong><br />

Myrtaceae 1<br />

Myrtaceae 2<br />

Myrtaceae 3<br />

Pimenta<br />

Craveiro Si Fo<br />

pseudocaryophyllus<br />

Blume<br />

28 Proteaceae Roupala brasiliensis Kl. Carvalho Pi, Si, St<br />

29 Rosaceae Prunus brasiliensis Schott Pessegueirobravo<br />

Si, St<br />

ex Spr.<br />

30 Rubiaceae Coussarea contracta pimenteira St<br />

(Walp.) Muell.Arg.<br />

Psychotria leiocarpa Gr<strong>and</strong>iúva St<br />

Cham. et S.<br />

31 Rutaceae Citrus lemon L. Limoeiro Ex R Fo<br />

Zanthoxyllum rhoifolium Mamica-de-<br />

Pi, Si, St<br />

cadela<br />

Chal-chal Pi, Si, St VC<br />

Cupania vernalis Camb. Cuvantã Pi, Si, St VC<br />

Matayba elaeagnoides Miguelpintado<br />

Pi, Si, St VC<br />

Radlk.<br />

33 Solanaceae Solanaceae 1 R<br />

34 Styracaceae Styrax leprosus Hook. et Carne-de-vaca Si, St NC Ti<br />

Arn.<br />

35 Verbenaceae Aegiphila sellowiana Tamanqueiro Pi, Si, St Ha<br />

Cham.<br />

Vitex megapotamica (Spr.) Tarumã St, Cl C Me<br />

Moldenke<br />

* PI= initial; Si = initial secondary; St = late secondary; Cl= climacic; Ex = non-native; ** VC- very common:<br />

DR≥5%; C- common: 1%>DR%


R.S. Moro & T.K. Pereira 2010. Evaluating an old sustainable national forest in south Brazil<br />

273<br />

b) including Araucaria, basal area ranges from 15 to 54m 2 /ha <strong>and</strong> dbh, from 3 to 62cm; c) liana<br />

<strong>and</strong> epiphytes are rare; there are few grasses <strong>and</strong> litter reaches more than 10 cm in the most part<br />

of the sampled sites; d) gaps could be common, sparsely located; e) there is no dominance of a<br />

few species over the others in the st<strong>and</strong>s. Typical species of this regeneration stadium are<br />

Casearia sylvestris, Matayba elaeagnoides, Capsicodendron dinisii, Eugenia uniflora,<br />

Dalbergia brasiliensis, Clethra scabra, Casearia lasiophylla, Alophyllus edulis, <strong>and</strong> Cupania<br />

vernalis. It was identified several endangered species as Dicksonia sellowiana (IBAMA 1992),<br />

Roupala brasiliensis, Apuleia leiocarpa, <strong>and</strong> Nect<strong>and</strong>ra megapotamica (Res. SEMA/IAP 31/98).<br />

There were endangered species <strong>and</strong> bioindicators that enhance the ecological importance of this<br />

National <strong>Forest</strong>. Thus, this one should not be exploited despite their original purpose <strong>and</strong><br />

deserves na integral conservation status.<br />

Acknowledgements: To ICMBio for the financial <strong>and</strong> field support; to Dr. Carlos Hugo<br />

Rocha for his l<strong>and</strong>scape´s components analysis disposal.<br />

References<br />

Castella, P.R. <strong>and</strong> Britez, R.M., 2004. A Floresta com Araucária no Paraná: conservação e<br />

diagnóstico dos remanescentes florestais.MMA, Brasília, 236p.<br />

CONAMA – Conselho Nacional de Meio Ambiente. Res. 2/94. Define vegetação primária e<br />

secundária nos estágios inicial, médio e avançado de regeneração da Mata Atlântica no<br />

Estado do Paraná.<br />

IBAMA - Instituto Brasileiro do Meio Ambiente e dos Recursos Naturais Renováveis. Portaria n.<br />

006/92-N/92. Apresenta a Lista Oficial de Espécies da Flora Brasileira Ameaçadas de<br />

Extinção.<br />

Shepherd, G.J., 1994. FITOPAC: manual do usuário. UNICAMP, Campinas, 25p.<br />

SNUC - Sistema Nacional de Unidades de Conservação da Natureza. Lei nº 9.985/2000.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & C.H Rocha 2010. A methodological proposal for restoration of forests in southern Brazil<br />

274<br />

A methodological proposal for restoration of forests in southern Brazil<br />

Rosemeri Segecin Moro 1* & Carlos Hugo Rocha 2<br />

1 DEBIO- Universidade Estadual de Ponta Grossa, PR, Brazil<br />

2 DESOLOS- Universidade Estadual de Ponta Grossa, PR, Brazil<br />

Abstract<br />

In order to comply with the Brazilian <strong>Forest</strong> Code, l<strong>and</strong>owners except in the Amazon must set<br />

aside at least 20% of their l<strong>and</strong> area as Legal Reserves. Regional projects developed a<br />

restoration model that is both ecologically <strong>and</strong> economically viable. By mixing native <strong>and</strong><br />

exotic (Eucalyptus) species, timber production is possible in 20 years. Adding to wood<br />

production, other benefits such as carbon offset, soil <strong>and</strong> water protection, increased<br />

biodiversity, <strong>and</strong> species conservation are ensured through this strategy. To enhance chances for<br />

adoption of such strategy on the highl<strong>and</strong>s in Southern Brazil, we propose to use the native<br />

conifer Araucaria instead of Eucalyptus in the system. A 30-year evaluation between<br />

reforestation <strong>and</strong> natural Araucaria canopies showed that biodiversity could be reached using<br />

planted forest as well setting natural areas aside. The presence of endangered species in both<br />

situations indicated the ecological importance of this alternative.<br />

Keywords: Araucaria forest – forest restoration – planted forests<br />

1. Subject<br />

Timber <strong>and</strong> cellulose production make up an expressive portion of the Brazilian<br />

economy. Planted forests are seen as a l<strong>and</strong> use form of nature conservation on the highl<strong>and</strong>s in<br />

southern Brazil. They are important sources of timber that would otherwise be exploited from<br />

the natural forests. In order to comply with the Brazilian <strong>Forest</strong> Code, l<strong>and</strong>owners must set aside<br />

at least 20% of their l<strong>and</strong> area as Legal Reserves (except in the Amazon, where at least 80%<br />

must be preserved) – l<strong>and</strong> properties that are short of natural forest cover to meet the minimum<br />

required Legal Reserve areas are required to restore it by either natural or artificial regeneration.<br />

The Paraná Biodiversity Project, with support from the State Government <strong>and</strong> the World<br />

Bank, developed a Legal Reserve restoration model that is both ecologically <strong>and</strong> economically<br />

viable. It is an effort to combine conservation goals with legal requirements <strong>and</strong> sustainable<br />

development. It is based on a forest management strategy on planted st<strong>and</strong>s with native tree<br />

species mixed with exotics (Eucalyptus). Timber production is planned on a 20-year span. After<br />

harvesting Eucalyptus species over that period, only native trees are left to start the restoration<br />

of the natural ecosystem. In addition to wood production, other benefits are ensured through the<br />

adoption of this strategy such as carbon offset, soil <strong>and</strong> water protection, increased biodiversity,<br />

<strong>and</strong> species conservation. All these factors contribute to increase forest fragment sizes <strong>and</strong><br />

l<strong>and</strong>scape connectivity. In order to increase the number of adepts to this strategy on the<br />

highl<strong>and</strong>s in Southern Brazil, we propose to use only native species, including Araucaria<br />

angustifolia instead of Eucalyptus as the major timber source in the system.<br />

* Corresponding author. Tel.:55-42-3223-4417 - Fax:55-42-3224-4747<br />

Email address: rsmoro@uepg.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & C.H Rocha 2010. A methodological proposal for restoration of forests in southern Brazil<br />

275<br />

2. Base for the proposal<br />

Surveys were carried out in 0.01-ha plots laid out in a 30-yr-old 400 ha planted Araucaria<br />

st<strong>and</strong> <strong>and</strong> a natural 320 ha Araucaria forest at Açungui National <strong>Forest</strong> in Campo Largo, State<br />

of Paraná, Southern Brazil (25 o 10'41"S - 25 o 14'18"W). The dominant soil types in the region are<br />

Inceptisol <strong>and</strong> Ultisol. The climate is classified as Cfb (under Koeppen’s climate classification<br />

system), at elevations between 640 m <strong>and</strong> 905 m above sea level. Species composition <strong>and</strong><br />

several forest structure parameters are summarized in Table 1.<br />

More than a half of the species found in both planted <strong>and</strong> natural Araucaria st<strong>and</strong>s were<br />

represented by Flacourtiaceae, Lauraceae <strong>and</strong> Myrtaceae. Sapindaceae <strong>and</strong> Euphorbiaceae were<br />

found only in the natural Araucaria st<strong>and</strong>. The most abundant species found under the planted<br />

Araucaria canopy were of Agavaceae, Flacourtiaceae, Sapindaceae, <strong>and</strong> Cyatheaceae families.<br />

Under the natural Araucaria canopy, the most abundant families were Sapindaceae, Myrtaceae,<br />

Canellaceae, <strong>and</strong> Euphorbiaceae. In terms of the Importance Value Index (IVI), the<br />

Flacourtiaceae, Agavaceae, Cyatheaceae, Sapindaceae, <strong>and</strong> Fabaceae were the most important<br />

families found under planted Araucaria st<strong>and</strong>s while Sapindaceae, Myrtaceae, Flacourtiaceae,<br />

<strong>and</strong> Araucariaceae were the most important families found under the natural forest canopy.<br />

Araucaria trees produced similar height <strong>and</strong> crown width in both planted <strong>and</strong> natural<br />

st<strong>and</strong>s. The species with the highest relative density under the planted Araucaria st<strong>and</strong> were<br />

Casearia sylvestris, Allophyllus edulis, Clethra scabra, Dalbergia brasiliensis, <strong>and</strong> Matayba<br />

elaeagnoides (see figure 1). At shrub level, the st<strong>and</strong> was dominated by species such as<br />

Cordyline dracaenoides, Cyathea corcovadensis, <strong>and</strong> Cyathea schanschin. The most frequent<br />

species were Cordyline dracaenoides, Cyathea corcovadensis, Matayba elaeagnoides, Casearia<br />

sylvestris, Casearia lasiophylla, Dalbergia brasiliensis, Clethra scabra, Cupania vernalis,<br />

Casearia obliqua, Cedrella fissilis, Allophyllus edulis, Cyathea schanschin, Guatteria australis,<br />

Casearia inaequilatera, Vitex megapotamica, Cabralea canjerana, <strong>and</strong> Myrcia rostrata.<br />

In the natural st<strong>and</strong> (see figure 2), only Araucaria angustifolia <strong>and</strong> Matayba<br />

elaeagnoides stood out over the canopy. The main species present in the st<strong>and</strong> were Matayba<br />

elaeagnoides, Casearia sylvestris, Allophyllus edulis, Capsicodendron dinisii, Araucaria<br />

angustifolia, <strong>and</strong> Eugenia uniflora. Among shrubs, the most frequent species were Mollinedia<br />

clavigera, Sebastiania brasiliensis, <strong>and</strong> Myrcia hatschbachii.<br />

Natural <strong>and</strong> planted Araucaria st<strong>and</strong>s were similar in regard to both Shannon index <strong>and</strong><br />

evenness. Simpson index indicated that there is no dominance of a few species over the others<br />

in either st<strong>and</strong>.<br />

3. Benefits<br />

Both natural <strong>and</strong> planted Araucaria st<strong>and</strong>s showed similar phytosociological parameters.<br />

Therefore, our suggestion is that planted Araucaria st<strong>and</strong>s are suitable to be used in place of<br />

Eucalyptus for conservation purposes.<br />

Although Araucaria is highly site dem<strong>and</strong>ing, when planted on good quality sites, it can<br />

produce as much as 50 m 3 /ha.yr. Commercial Araucaria timber volume production was higher<br />

in the planted st<strong>and</strong>.<br />

Eighteen-year-old planted Araucaria st<strong>and</strong>s have shown high efficiency in conserving<br />

soil organic carbon (Guedes 2005). Thus, wood stock was maintained in levels equivalent to<br />

those in natural forests on good sites. Soil organic carbon was maintained on the site at rates<br />

ranging from 23 to 56 g/kg, mostly as litter. In a 15-yr-old planted Araucaria st<strong>and</strong>, annual dry<br />

matter accumulation in the litter varied from 5.0 to 6.4 Mg/ha (Koehler et al. 1987) while, in<br />

natural forests, it could reach 5.9 to 8.3 Mg/ha (Fern<strong>and</strong>es <strong>and</strong> Backes 1998; Floss et al. 1999;<br />

Figueiredo Filho et al. 2003).<br />

The presence of endangered species in both types of Araucaria forests indicated the<br />

ecological importance of the low density st<strong>and</strong> as a suitable alternative for the purpose of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & C.H Rocha 2010. A methodological proposal for restoration of forests in southern Brazil<br />

276<br />

combined timber production <strong>and</strong> natural forest conservation. We propose a management option<br />

involving an initial st<strong>and</strong> density of 750 trees/ha with at least one thinning down to 250 trees/ha<br />

at 15 years of age.<br />

Table 1: Tree species <strong>and</strong> families found under the canopies of both planted <strong>and</strong> natural Araucaria<br />

angustifolia st<strong>and</strong>s at Açungui National <strong>Forest</strong> in Campo Largo (Paraná-Brazil).<br />

Under planted Araucaria<br />

canopy<br />

Under natural Araucaria<br />

canopy<br />

Sampled area (ha) 0.12 -<br />

Number of tree species 61 53<br />

Number of families 30 19<br />

Density (plant/ha) 550 1,680<br />

Richest families (number of species) Flacourtiaceae (7)<br />

Lauraceae (5)<br />

Myrtaceae (4)<br />

Abundant families (count) Araucariaceae (66)<br />

Agavaceae (67)<br />

Flacourtiaceae (63)<br />

Sapindaceae (42)<br />

Cyatheaceae (25)<br />

Families with highest IVI (%) Araucariaceae (29.5)<br />

Flacourtiaceae (10.3)<br />

Agavaceae (9.2)<br />

Cyatheaceae (7.8)<br />

Sapindaceae (7.1)<br />

dead (4.4)<br />

Myrtaceae (12)<br />

Flacourtiaceae (5)<br />

Lauraceae (4)<br />

Sapindaceae (2)<br />

Euphorbiaceae (2)<br />

Sapindaceae (26)<br />

Myrtaceae (23)<br />

Canellaceae (8)<br />

Euphorbiaceae (8)<br />

Sapindaceae (21.9)<br />

dead (12.5)<br />

Myrtaceae (12.4)<br />

Flacourtiaceae (10.5)<br />

Araucariaceae (9.6)<br />

Fabaceae (3.8)<br />

Araucaria – mean height (m) 15 16<br />

Araucaria - mean Dbh (cm) 25 29<br />

Canopy – mean height (m) 7.1 (single level) 11.0 <strong>and</strong> 5.0 (two levels)<br />

Canopy – mean Dbh (cm) 11.2 11.0<br />

Shrub height range (m) 3.0-5.3 3.0-5.3<br />

Species with the highest relative density<br />

(in decreasing order)<br />

Shrub species<br />

Araucaria angustifolia<br />

Casearia sylvestris<br />

Allophyllus edulis<br />

Clethra scabra<br />

Dalbergia brasiliensis<br />

Matayba elaeagnoides.<br />

Cordyline dracaenoides<br />

Cyathea corcovadensis,<br />

Cyathea schanschin<br />

Shannon-Wienner Index 3.15 3.43<br />

Evenness 0.77 0.86<br />

Simpson Index 0.93 0.95<br />

Matayba elaeagnoides<br />

Casearia sylvestris<br />

Allophyllus edulis<br />

Capsicodendron dinisii<br />

Araucaria angustifolia<br />

Eugenia uniflora.<br />

Mollinedia clavigera<br />

Sebastiania brasiliensis<br />

Myrcia hatschbachii<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & C.H Rocha 2010. A methodological proposal for restoration of forests in southern Brazil<br />

277<br />

Figure 1: Diagramatic profile (5 X 40 m) of 30-year-old planted Araucaria st<strong>and</strong> at Açungui National<br />

<strong>Forest</strong> in Campo Largo (Paraná-Brazil). 1. Araucaria angustifolia; 2. Cordyline dracaenoides; 3.<br />

Cyathea corcovadensis; 4. Casearia sylvestris; 5. Allophyllus edulis; 6. Clethra scabra; 7. Dalbergia<br />

brasiliensis; 8. Matayba elaeagnoides; 9. Casearia lasiophylla; 10. Cupania vernalis. Author: R. F. Moro.<br />

Fig. 2: Diagramatic profile (5 X 20 m) of a 30-year-old natural Araucaria st<strong>and</strong> at Açungui National<br />

<strong>Forest</strong> in Campo Largo (Paraná-Brazil). 1. Araucaria angustifolia; 2. Matayba elaeagnoides; 3. Casearia<br />

sylvestris; 4. Capsicodendron dinisii; 5. Allophyllus edulis; 6. Eugenia uniflora; 7. Mollinedia clavigera;<br />

8. Sebastiania brasiliensis. Author: R. F. Moro.<br />

Acknowledgements: To ICMBio for the financial <strong>and</strong> field support; to Dr. Jarbas Shimizu for<br />

the technical discussions.<br />

References<br />

Fern<strong>and</strong>es, A.V. <strong>and</strong> Backes, A., 1998. Produtividade primária em floresta com Araucaria<br />

angustifolia no Rio Gr<strong>and</strong>e do Sul. Iheringia, Série Botânica, 50: 63-78.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.S. Moro & C.H Rocha 2010. A methodological proposal for restoration of forests in southern Brazil<br />

278<br />

Figueiredo Filho, A.; Moraes, G.F.; Schaaf, L.B. <strong>and</strong> Figueiredo, D.J., 2003. Avaliação<br />

estacional da deposição de serrapilheira em uma Floresta Ombrófila Mista localizada no<br />

sul do estado do Paraná. Ciência Florestal, 13: 11-18.<br />

Floss, P.A.; Caldato, S.L. <strong>and</strong> Bohner, J.A.M., 1999. Produção e decomposição de serapilheira<br />

na Floresta Ombrófila Mista da Reserva Florestal da EPAGRI/EMBRAPA de Caçador,<br />

SC. Agropecuária Catarinense, 12 (2): 19-22.<br />

Guedes, S.F.F., 2005. Carbono orgânico e atributos químicos do solo em áreas florestais no<br />

Planalto dos Campos Gerais, SC. UDESC, Lages, 58 p.<br />

Koehler, C.W.; Reissmann, C.B.; <strong>and</strong> Koeler, H.S., 1987. Deposição de resíduos orgânicos<br />

(serapilheira) e nutrientes em plantios de Araucaria angustifolia em função do sítio. Rev.<br />

Setor Ciênc. Agrárias, 9: 89-96.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Núñez et al. 2010. L<strong>and</strong>scape integration of Mediterranean reforestations: identification of best practices in Madrid region<br />

279<br />

L<strong>and</strong>scape integration of Mediterranean reforestations: identification<br />

of best practices in Madrid region<br />

Victoria Núñez 1* , María Dolores Velarde 2 & Antonio García-Abril 1<br />

1<br />

Technical University of Madrid, Research Group for Sustainable Management, E.T.S.I.<br />

Montes, Spain<br />

2<br />

School of Experimental <strong>and</strong> Technical Sciences, Rey Juan Carlos University of<br />

Madrid, Spain<br />

Abstract<br />

The European L<strong>and</strong>scape Convention was approved in 2000 by The European Council with the<br />

aim of protecting, managing <strong>and</strong> planning l<strong>and</strong>scapes in Europe. In 2008, this Convention<br />

entered into force in Spain, so nowadays it is m<strong>and</strong>atory to develop specific or sectorial<br />

strategies.<br />

We propose a methodology to define criteria <strong>and</strong> identify best practices for l<strong>and</strong>scape<br />

integration of reforestations in the Mediterranean region, based on l<strong>and</strong>scape ecology <strong>and</strong> visual<br />

properties. The methodology encompasses the following phases: i) Identification of the main<br />

relationships of l<strong>and</strong>scape with biodiversity, connectivity <strong>and</strong> public preferences; ii) synthesis of<br />

principles <strong>and</strong> criteria for l<strong>and</strong>scape integration of Mediterranean reforestations; iii) Definition<br />

of detailed l<strong>and</strong>scape criteria, both qualitative <strong>and</strong> quantitative, for reforestation design, species<br />

selection, soil preparation <strong>and</strong> protection of the reforestation area.<br />

We applied this methodology to reforestations in Madrid Region. As a result we typified 30 best<br />

practices that may be of use for reforestation in the Mediterranean region.<br />

Keywords: L<strong>and</strong>scape, reforestation, integration, Mediterranean region, best practices<br />

1. Introduction<br />

The importance of l<strong>and</strong>scape in all aspects of forest planning <strong>and</strong> management has recently<br />

become significant due to a number of factors, such as the Rio-Helsinki process, the<br />

requirements of certification <strong>and</strong> an international movement favouring more natural forest<br />

management.<br />

Moreover, the European L<strong>and</strong>scape Convention approved in 2000 by The European Council<br />

with the aim of protecting, managing <strong>and</strong> planning l<strong>and</strong>scapes is m<strong>and</strong>atory nowadays in Spain.<br />

So, specific or sectorial strategies must be developed.<br />

All these factors began to change thinking about appropriate silvicultural systems for forests. So<br />

comprehensive l<strong>and</strong>scape plans are necessary when new planting is undertaken on a substantial<br />

scale.<br />

2. Methodology<br />

We developed a methodology to define criteria <strong>and</strong> identify best practices for l<strong>and</strong>scape<br />

integration of reforestations in the Mediterranean region, based on l<strong>and</strong>scape ecology (Farina<br />

2006) <strong>and</strong> visual properties. The methodology (Figure 1) encompasses the following phases: i)<br />

Identification of the main relationships of l<strong>and</strong>scape with biodiversity, connectivity <strong>and</strong> public<br />

* Corresponding author. Tel.: +34913366401 - Fax: +34915439557<br />

Email address: mvn@iies.es<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Núñez et al. 2010. L<strong>and</strong>scape integration of Mediterranean reforestations: identification of best practices in Madrid region<br />

280<br />

preferences; ii) synthesis of principles <strong>and</strong> criteria for l<strong>and</strong>scape integration of Mediterranean<br />

reforestations; iii) Definition of detailed l<strong>and</strong>scape criteria, both qualitative <strong>and</strong> quantitative, for<br />

reforestation design, species selection, soil preparation <strong>and</strong> protection of the reforestation area.<br />

3. Results<br />

We identified the following list items that must be taken into account for new mediterranean<br />

reforestation design. Underst<strong>and</strong>ing these factors, the interactions between them, <strong>and</strong> how they<br />

are influenced by st<strong>and</strong> management is essential in order to provide guidance to reforestation<br />

design under visual <strong>and</strong> l<strong>and</strong>scape ecology criteria.<br />

We applied this methodology to Madrid Region reforestations. Field visits were made<br />

throughout the forests <strong>and</strong> managers were interviewed. As a result we typified 30 best practices<br />

that may be of use for reforestation in the Mediterranean region.<br />

3.1. Main relationships of forest l<strong>and</strong>scape with biodiversity, connectivity <strong>and</strong> public<br />

preferences.<br />

1. <strong>Forest</strong> l<strong>and</strong>scape <strong>and</strong> biodiversity<br />

Mosaic structure<br />

St<strong>and</strong> structure<br />

Deadwood <strong>and</strong> large trees<br />

Minimum dynamic area (Pickett <strong>and</strong> Thompson 1978) or independent<br />

forest size<br />

2. <strong>Forest</strong> l<strong>and</strong>scape <strong>and</strong> connectivity:<br />

Size of isolated patches<br />

Edges geometry <strong>and</strong> size<br />

Matrix <strong>and</strong> patches pattern<br />

3. <strong>Forest</strong> l<strong>and</strong>scape <strong>and</strong> public preferences<br />

Aesthetic preferences<br />

Time factor<br />

Spatial activities distribution<br />

3.2. Synthesis of principles <strong>and</strong> criteria for l<strong>and</strong>scape integration of reforestations in<br />

the Mediterranean region<br />

Principles<br />

1. Multiscale approach to reforestation design<br />

2. Try to match up aesthetic <strong>and</strong> ecological criteria<br />

3. Take into account public preferences<br />

4. Try to simulate nature when designing reforestation<br />

5. Mask inevitable negative l<strong>and</strong>scape impacts<br />

Criteria for external l<strong>and</strong>scape: broad-scale approach<br />

1. Avoid fragmentation <strong>and</strong> improve ecosystem connectivity<br />

2. Increase biodiversity by species introduction<br />

3. Mitigate the visual impacts of reforestation boundary<br />

4. Adapt lineal structures to the terrain<br />

5. Fit reforestation activities to l<strong>and</strong>scape scale<br />

Criteria for internal l<strong>and</strong>scape: small-scale approach<br />

1. Protect riversides <strong>and</strong> shores<br />

2. Design reforestation edges with gradual slope<br />

3. Preserve or create gaps in forest cover<br />

4. Preserve large <strong>and</strong> dead trees<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Núñez et al. 2010. L<strong>and</strong>scape integration of Mediterranean reforestations: identification of best practices in Madrid region<br />

281<br />

5. Try to integrate infrastructures <strong>and</strong> equipment<br />

6. Do not disturb the genius loci (spirit of the place)<br />

3.3 Detailed l<strong>and</strong>scape criteria<br />

1. Reforestation pattern<br />

2. Reforestation size<br />

3. Planting density<br />

4. Species selection<br />

5. Soil preparation<br />

6. Protection of the reforestation area<br />

7. Initial pruning <strong>and</strong> thinning treatments.<br />

We only present here some of the main detailed l<strong>and</strong>scape criteria as they exceed the length of<br />

this communication.<br />

3.3.1. Reforestation pattern<br />

We assessed different patterns depending on the reforestation objective (table 1). A mosaic of<br />

structures is most proposed because it involves biodiversity increase (Díaz Pineda <strong>and</strong> Schmidtz<br />

2003; Farina 2006; de Zavala et al. 2008), stability improvement (Margaleff 1993), connectivity<br />

strength (Pino et al. 2000; De Lucio et al. 2003) <strong>and</strong> enhanced visual quality (Ammer <strong>and</strong><br />

Pröbstl 1991).<br />

3.3.2. Reforestation size<br />

Mediterranean l<strong>and</strong>scape reforestations must include forested patches of different size. But, for<br />

inside-forest species presence, patches over 100 ha must be part of the mosaic (Dajoz 2002). If<br />

pine reforestation is adopted, patches over 70 ha are needed to mature all forest phases<br />

development (Korpel 1982).<br />

3.3.3. Planting density<br />

Planting density recommendations depend on reforestation objective: wood production (1000-<br />

2000 trees/ha), erosion protection (3000-2000 trees/ha), natural forest reconstruction (


V. Núñez et al. 2010. L<strong>and</strong>scape integration of Mediterranean reforestations: identification of best practices in Madrid region<br />

282<br />

The punctual soil-preparation techniques making a hole <strong>and</strong> constructing small runoff collectors<br />

(minicatchments) are preferred to linear <strong>and</strong> surface treatments that prepare the soil in general<br />

all over the plantation surface.<br />

Areas that have normally been considered more difficult to restore due to a de-structured profile,<br />

to the presence of superficial physical crusts <strong>and</strong> to a scarce vegetal cover are the ones showing<br />

a better response to minicatchments system (Bocio et al. 2004; De Simón 2006). Best practices<br />

were identified in “Navas Pinarejo” forest <strong>and</strong> “Sureste” Regional Park.<br />

3.3.6. Protection of the reforestation area<br />

Individual shelters do not show consistent better results than boundary fences for reforestation<br />

survival (Costa 2003; Sharew <strong>and</strong> Hairston-Strang 2005; Bergez <strong>and</strong> Dupraz 2009). Visual<br />

impact depends on the colour integration <strong>and</strong> visibility (Ammer <strong>and</strong> Pröbstl 1991). They cause<br />

temporal impact if they are removed. Best practices were described in “Canencia”, “Navas y<br />

Pinarejo” <strong>and</strong> “La Morcuera” forests.<br />

4. Conclusion<br />

The proposed array of principles <strong>and</strong> criteria offers considerable promise to l<strong>and</strong>scape discipline<br />

incorporation in reforestation activities <strong>and</strong> should be view as a push to sustainable forest<br />

management <strong>and</strong> best practices identification.<br />

References<br />

Ammer, U. <strong>and</strong> Pröbstl, U., 1991. Freizeit und Natur. Probleme und Lösungsmöglichkeiten<br />

einer ökologisch verträglichen Freizaitnutzung (Tiempo Libre y Naturaleza. Problemas<br />

y soluciones de un uso ecológico y sostenible del recreo). Verlag Paul Parey, Hamburg.<br />

Bergez, J.E. <strong>and</strong> Dupraz, C., 2009. Radiation <strong>and</strong> thermal microclimate in tree shelter. Agric.<br />

For. Meteorol., 149 (1): 179-186.<br />

Bocio, I., Navarro, F.B., Ripoll, M.A., Jiménez, M.N. <strong>and</strong> De Simón, E., 2004. Holm oak<br />

(Quercus rotundifolia Lam.) <strong>and</strong> Aleppo pine (Pinus halepensis Mill.) response to<br />

different soil preparation techniques applied to forestation in ab<strong>and</strong>oned farml<strong>and</strong>.<br />

Annals of <strong>Forest</strong> Science, 61 (2): 171-178.<br />

Cortina, J., Peñuelas, J.L., Puértolas, J., Savé, R. <strong>and</strong> Vilagrosa, A., 2006. Calidad de planta<br />

forestal para la restauración en ambientes mediterráneos. Estado actual de<br />

conocimientos. Organismo Autónomo Parques Nacionales. Ministerio de Medio<br />

Ambiente, Madrid.<br />

Costa, J.C., 2003. Evaluación de la aplicación de tubos y mejoradores en repoblaciones<br />

forestales. Dirección General de Gestión del Medio Natural, Consejería de Medio<br />

Ambiente, Junta de Andalucía, Sevilla.<br />

Dajoz, R., 2002. Tratado de ecología. Mundi Prensa, Madrid.<br />

De Lucio, J.V., Atauri Mezquida, J.A., Sastre Olmos, P. <strong>and</strong> Martínez Al<strong>and</strong>i, C., 2003.<br />

Conectividad y redes de espacios naturales protegidos. Del modelo teórico a la visión<br />

práctica de la gestión. In: García Mora, R.C. (Ed), Conectividad ambiental: las áreas<br />

protegidas en la cuenca mediterránea, Sevilla, Consejería de Medio Ambiente. Junta<br />

de Andalucía, pp. 29 – 53.<br />

De Simón, E., Ripoll, M.A., Fernández, E., Navarro, F.B., Jiménez, M.N., Gallego, E., 2006.<br />

Eficacia de las microcuencas en la supervivencia del pino carrasco (Pinus halepensis<br />

Mill.) y de la encina (Quercus ilex L. subsp. ballota (Desf.) Samp.) en distintos<br />

ambientes mediterráneos. Investigación agraria: Sistemas de recursos forestales, 15<br />

(2): 218-230.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Núñez et al. 2010. L<strong>and</strong>scape integration of Mediterranean reforestations: identification of best practices in Madrid region<br />

283<br />

de Zavala, M.A., Zamora, R., Pulido, F., Blanco, J.A., Imbert, B., Marañón, T., Castillo, F. <strong>and</strong><br />

Valladares, F., 2008. Nuevas perspectivas en la conservación, restauración y gestión<br />

sostenible del bosque mediterráneo. In: Valladares, F.E. (Ed), Ecología del bosque<br />

mediterráneo en un mundo cambiante, Madrid, Ministerio de Medio Ambiente.<br />

Organismo autónomo de Parques Nacionales, pp. 511-532.<br />

Díaz Pineda, F. <strong>and</strong> Schmidtz, F.M., 2003. Tramas espaciales del paisaje. Conceptos,<br />

aplicabilidad y temas urgentes para la planificación territorial. In: García Mora, M.R.C.<br />

(Ed), Conectividad ambiental: las áreas protegidas en la Cuenca Mediterránea,<br />

Sevilla, Consejería de Medio Ambiente Junta de Andalucía, pp. 9-28.<br />

Farina, A., 2006. Principles <strong>and</strong> methods in l<strong>and</strong>scape ecology: toward a science of l<strong>and</strong>scape.<br />

Springer, Dordrecht.<br />

Hernández, C.G., 1998. Ecología y fisiología de la Dehesa. In: Díaz-Ambrona, C.H. (Ed), La<br />

Dehesa. Aprovechamiento sostenible de los recursos naturales, Editorial Agrícola<br />

Española S.A., pp. 53-94.<br />

Korpel, S., 1982. Degreee of equilibrium <strong>and</strong> dinamical changes of forest on example of natural<br />

forest of slovakia. Acta facultatis forestalis zvolen, XXIV.<br />

Margaleff, R., 1993. Teoría de los sistemas ecológicos. Universidad de Barcelona.<br />

Pemán, J., Navarro, R. <strong>and</strong> Serrada, R., 2006. Elección de especies en las repoblaciones<br />

forestales. Contribuciones del profesor Ruiz de la Torre. Investigación agraria:<br />

Sistemas de recursos forestales, 15 (1): 87-102.<br />

Pickett, S.T.A. <strong>and</strong> Thompson, J.N., 1978. Patch dynamics <strong>and</strong> the design of nature reserves.<br />

Biological Conservation, 13 (1): 27-37.<br />

Pino, J., Roda, F., Ribas, J. <strong>and</strong> Pons, X., 2000. L<strong>and</strong>scape structure <strong>and</strong> bird species richness:<br />

implications for conservation in rural areas between natural parks. L<strong>and</strong>scape <strong>and</strong><br />

Urban Planning, 49 (1-2): 35-48.<br />

Serrada, R., 2009. Repoblaciones forestales: concepto y planificación.<br />

http://www.forestales.upm.es/Descarga.aspxcont=unidad&sec=14&id=139, Madrid.<br />

Serrada, R., Montero, G. <strong>and</strong> Reque, J.A., 2008. Compendio de Selvicultura Aplicada en<br />

España. Madrid, I.N.I.A.<br />

Sharew, H. <strong>and</strong> Hairston-Strang, A., 2005. A comparison of seedling growth <strong>and</strong> light<br />

transmission among tree shelters. North. J. Appl. For., 22 (2): 102-110.<br />

Figures <strong>and</strong> tables<br />

Figure 1: Methodology<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Núñez et al. 2010. L<strong>and</strong>scape integration of Mediterranean reforestations: identification of best practices in Madrid region<br />

284<br />

L<strong>and</strong> cover class<br />

Erosion in<br />

grassl<strong>and</strong>, shrubl<strong>and</strong><br />

or scarce tree<br />

density forest<br />

Grassl<strong>and</strong>, no<br />

erosion risk<br />

Shrub l<strong>and</strong>, no<br />

erosion signs, poor<br />

soil<br />

Shrub l<strong>and</strong>, no<br />

erosion signs, deep<br />

soil<br />

Gaps in forest cover<br />

with<br />

grassl<strong>and</strong>/shrubl<strong>and</strong>,<br />

no erosion<br />

<strong>Forest</strong> with scarce<br />

canopy cover<br />

<strong>Forest</strong> with dense<br />

canopy cover<br />

Table 1: Reforestation patterns<br />

Main reforestation objective<br />

Wood production Erosion control Natural <strong>Forest</strong><br />

---<br />

Preserve grassl<strong>and</strong><br />

gaps to allow fauna<br />

movements.<br />

Incorporate auxiliary<br />

species for soil<br />

improvement<br />

Reforestation not<br />

recommended<br />

Preserve shrubl<strong>and</strong><br />

gaps to allow fauna<br />

movements.<br />

Incorporate auxiliary<br />

species for soil<br />

improvement<br />

Small gap:<br />

reforestation not<br />

recommended.<br />

Big gap:<br />

Plurispecific non<br />

continuous<br />

reforestation<br />

Incorporate auxiliary<br />

species for soil<br />

improvement <strong>and</strong><br />

future wood<br />

production<br />

Use felling gaps for<br />

plurispecific<br />

reforestation of<br />

individuals or by<br />

small groups.<br />

Incorporate auxiliary<br />

species for soil<br />

improvement <strong>and</strong><br />

future wood<br />

production<br />

Plurispecific<br />

continuous<br />

reforestation,<br />

including shrub<br />

<strong>and</strong> tree species<br />

---<br />

---<br />

---<br />

---<br />

---<br />

---<br />

Burned areas<br />

restoration<br />

--- ---<br />

Plurispecific non<br />

continuous<br />

reforestation. Wide<br />

grassl<strong>and</strong> gaps to<br />

preserve biotope.<br />

Mosaic pattern<br />

Reforestation not<br />

recommended<br />

Plurispecific non<br />

continuous<br />

reforestation.<br />

Preserve wide<br />

shrubl<strong>and</strong> gaps for<br />

biotope uses<br />

Small gap:<br />

reforestation not<br />

recommended. Big<br />

gap: Plurispecific<br />

reforestation of<br />

individuals or by<br />

small groups to<br />

improve biodiversity<br />

Plurispecific<br />

reforestation of<br />

individuals or by<br />

small groups. Growth<br />

must be allowed by<br />

existing trees<br />

Use natural cover<br />

gaps for plurispecific<br />

reforestation of<br />

individuals or by<br />

small groups.<br />

---<br />

Reforestation not<br />

recommended<br />

Plurispecific non<br />

continuous<br />

reforestation where<br />

natural regeneration<br />

does not occur.<br />

Preserve wide<br />

shrubl<strong>and</strong> gaps for<br />

biotope uses.<br />

Species enrichment<br />

where natural<br />

regeneration<br />

happens<br />

---<br />

---<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.G. Plexida & A.I. Sfougaris 2010. Conservation of priority bird species in a protected area of central Greece<br />

285<br />

Conservation of priority bird species in a protected area of central<br />

Greece using geographical location <strong>and</strong> GIS<br />

Sofia G. Plexida 1* & Athanassios I. Sfougaris 2<br />

Laboratory of Ecosystem <strong>and</strong> Biodiversity Management, Department of Agriculture,<br />

Crop Production <strong>and</strong> Rural Environment, University of Thessaly,<br />

Fytokou str., Ν. Ionia, 384 46 Volos, Greece<br />

Abstract<br />

The effects of habitat structure <strong>and</strong> spatial variation on bird community composition were<br />

studied in the protected area “Antichasia-Meteora mountains” (GR1440003), 82.635 km 2 . The<br />

census of bird diversity was conducted from late April until mid June, <strong>and</strong> in October 2008. In<br />

185 r<strong>and</strong>omly selected sampling plots, the following parameters were recorded: (i) habitat type;<br />

(ii) habitat cover variables <strong>and</strong> (iii) spatial variables. Stepwise regression analysis was used to<br />

investigate the relationships between bird species richness <strong>and</strong> vegetation. Geostatistics were<br />

used to examine the spatial distribution of priority species by semivariography followed by<br />

kriging interpolation. The results showed that bird species richness is correlated significantly<br />

with fallow l<strong>and</strong> (β=0.15, P50cm tall) (P


S.G. Plexida & A.I. Sfougaris 2010. Conservation of priority bird species in a protected area of central Greece<br />

286<br />

variables that regulate bird distrubution <strong>and</strong> (3) to analyse the spatial distribution of bird<br />

diversity, referring to richness <strong>and</strong> abundance.<br />

2. Methodology<br />

The Special Protected Area (SPA) “Mountain Antichasia – Meteora” studied (82.635ha), is<br />

extended between 250 to 1,400 m <strong>and</strong> is considered very important for its wild fauna (Meliadis<br />

et al. 2009). There are a few hedgerows among fields, while the main shrub species, Pyrus<br />

amygdaloformis; Quercus coccifera; Carpinus orientalis; Cotinus coggygria, occupy only 25%<br />

of the total area. Trees, mostly Q. pubescens, Q. frainetto <strong>and</strong> Q. cerris, cover 65% of the area.<br />

The rest 10% of the study area is covered by cultivations.<br />

2.1 Vegetation survey<br />

Vegetation measurements were taken within the same circular plots of 50m radius where the<br />

birds were sampled. Within each sample unit, five vegetation variables were measured: cover<br />

(%) of herbaceous plants, low shrubs, tall shrubs, trees, <strong>and</strong> cover (%) of bare ground (Table 1).<br />

In order to avoid observer-related biases in vegetation sampling (Prodon <strong>and</strong> Lebreton 1981), all<br />

vegetation parameter estimations were conducted by the same observer to control for interobserver<br />

variability (Morrison et al. 1992).<br />

2.2 Bird counts<br />

Bird counts took place in spring <strong>and</strong> late autumn 2008 on 185 plots using the point count<br />

method of 50m radius (0.785ha) (Ralph et al. 1995; Bibby et al. 2000). Each point was<br />

separated by at least 250m from all other points to minimize the probability of sampling the<br />

same bird more than once. All bird species using a plot were counted. These included all birds<br />

recorded on the ground <strong>and</strong> birds hunting over sample plots such as kestrels. Counts were made<br />

with binoculars Nikon 7218 Action 10x50mm by two observers simultaneously (Bibby et al.<br />

1992). The first set of counts were conducted to sample spring migrants that use this area as<br />

stopover site during their spring migration, <strong>and</strong> breeding birds, while the second one aimed at<br />

censusing the resident birds.<br />

2.3 Statistical analysis<br />

The overall effect of habitat characteristics (habitat type, habitat cover variables) was assessed<br />

by stepwise multiple regression with the variables recorded for each sample plot. In these<br />

analyses data were log(x+1) transformed to ensure normality. To explain a significant<br />

proportion of variation in field use for each bird species, we used stepwise regression analyses.<br />

The role of geographical location was explored by the ordinary kriging geostatistical models<br />

(Johnston et al. 2001). The “spatial” data matrix was conducted from x <strong>and</strong> y geographic<br />

coordinates. Non-normal variables were logarithmically transformed (Sokal <strong>and</strong> Rohlf 1981).<br />

3. Results<br />

The five vegetation variables measured in the 185 studied plots were not highly inter-correlated<br />

with the exception of shrubs presence (Table 2). There were two extremely significant variables,<br />

the low shrub cover <strong>and</strong> the tree cover (r= -0.698, P


S.G. Plexida & A.I. Sfougaris 2010. Conservation of priority bird species in a protected area of central Greece<br />

287<br />

Alauda arvensis (N=119), Turdus philomelos (N=106) <strong>and</strong> Lullula arborea (N=95). According<br />

to boxplots, bird richness <strong>and</strong> abundance varies more in farml<strong>and</strong> <strong>and</strong> fallowl<strong>and</strong> than the other<br />

habitat types (Figures 1 <strong>and</strong> 2). Specifically, stepwise multiple regression showed that bird<br />

richness is correlated significantly with fallow l<strong>and</strong> (β=0.15, P50cm tall) (P


S.G. Plexida & A.I. Sfougaris 2010. Conservation of priority bird species in a protected area of central Greece<br />

288<br />

Johnston, K., Ver Hoef J.M., Krivoruchko, K. <strong>and</strong> Lucas, N., 2001. Using ArcGIS geostatistical<br />

analyst. GIS by ESRT. ESRI, Redl<strong>and</strong>s, California.<br />

Johnson, G. <strong>and</strong> Freedman, B., 2002. Breeding birds in forestry plantations <strong>and</strong> natural forest in<br />

the vicinity of Fundy National Park, New Brunswick. Can. Field Nat., 116: 475–486.<br />

Maes, D., Gilbert, M., Titeux, N., Goffart, P. <strong>and</strong> Dennis, R.L., 2003. Prediction of butterfly<br />

diversity hotspots in Belgium: a comparison of statistically focused <strong>and</strong> l<strong>and</strong> usefocused<br />

models. Journal of Biogeography, 30: 1907-1920.<br />

Meliadis, I., Platis, P., Ainalis, A. <strong>and</strong> Meliadis, M., 2009. Monitoring <strong>and</strong> analysis of natural<br />

vegetation in a Special Protected Area of Mountain Antichasia—Meteora, central<br />

Greece. Environ Monit Assess., Vol. 163: 455-465.<br />

Morrison, M.L., Marcot, B.G. <strong>and</strong> Mannan, R.W., 1992. Wildlife–Habitat Relationships.<br />

Concepts <strong>and</strong> Applications. University of Wisconsin press, Wisconsin.<br />

Prodon, R. <strong>and</strong> Lebreton, J.D., 1981. Breeding avifauna of a Mediterranean succession: the<br />

holm <strong>and</strong> cork oak series in eastern Pyrenees. I: analyses <strong>and</strong> modelling of the structure<br />

gradient. Oikos, 37: 21–38.<br />

Ralph, C.J., Sauer, J.R. <strong>and</strong> Droege, S., 1995. Monitoring Bird Populations by Point Counts.<br />

General Technical Report. PSWGTR-149. Pacific Southwest Research Station, <strong>Forest</strong><br />

Service, U.S. Department of Agriculture, Albany, CA.<br />

S<strong>and</strong>ström, U.G., Angelstam, P. <strong>and</strong> Mikusinski, G., 2006. Ecological diversity of birds in<br />

relation to the structure of urban green space. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 77: 39-<br />

53.<br />

Santos, T., Tellería, J.L. <strong>and</strong> Carbonell, R., 2002. Bird conservation in fragmented<br />

Mediterranean forests in Spain: effects of geographical location, habitat <strong>and</strong> l<strong>and</strong>scape<br />

degradation. Biological Conservation, 105: 113-125.<br />

Scozzafava, S. <strong>and</strong> De Sanctis, A., 2006. Exploring the effects of l<strong>and</strong> ab<strong>and</strong>onment on<br />

habitat structures <strong>and</strong> on habitat suitability for three passerine species in a<br />

highl<strong>and</strong> area of Central Italy. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 75: 23-33.<br />

Sokal, R.R. <strong>and</strong> Rohlf, F.J., 1981. Biometry, 2 nd Edition. Freeman, New York.<br />

Tucker, G.M. <strong>and</strong> Evans, M.I., 1997. Habitats for Birds in Europe: A Conservation Strategy for<br />

the Wider Environment. BirdLife Conservation Series no. 6., BirdLife International,<br />

Cambridge.<br />

Table 1: Variables measured to characterise vegetation structure <strong>and</strong> tree <strong>and</strong> shrub composition in the<br />

sample plots 1 .<br />

1. HERBCOVER: % cover of herbaceous plants<br />

2. LSHCOVER: % cover of low shrubs<br />

3. TSHCOVER: % cover of tall shrubs<br />

4. TREECOVER: % cover of trees<br />

5. GRCOVER: % cover of bare ground<br />

1 Cover variables were recorded as percentages of the sample plot.<br />

Table 2: Pearson correlation coefficients between the five sampled vegetation variables in the studied<br />

plots (n=185).<br />

Herbaceous<br />

cover<br />

Low shrub<br />

cover<br />

Tall shrub<br />

cover<br />

Tree cover Bare ground<br />

cover<br />

Herbaceous cover 1 -0.418** -0.319** -0.698** -0.254**<br />

Low shrub cover 1 0.630** -0.161** 0.110<br />

Tall shrub cover 1 -0.250** -0.011<br />

Tree cover 1 -0.138<br />

Bare ground<br />

1<br />

cover<br />

P < 0.01, correlation is significant at the 0.01 level<br />

P < 0.001, correlation is significant at the 0.05 level<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.G. Plexida & A.I. Sfougaris 2010. Conservation of priority bird species in a protected area of central Greece<br />

289<br />

Table 3: List of the priority bird species in the entire sample area.<br />

Scientific name Common name Phenology 1 SPEC 2<br />

Category<br />

(2004)<br />

Birds<br />

Directive<br />

79/409<br />

Alauda arvensis Eurasian Skylark R 3 II/2<br />

Circaetus gallicus Short-toed eagle M 3 I<br />

Corvus corone Hooded Crow R II/2<br />

Corvus monedula Western Jackdaw R II/2<br />

Dendrocopos<br />

Middle Spotted<br />

R<br />

I<br />

medius<br />

Woodpecker<br />

Dendrocopos<br />

Syrian Woodpecker R I<br />

syriacus<br />

Garrulus<br />

Eurasian Jay R II/2<br />

gl<strong>and</strong>arius<br />

Falco eleonorae Eleonora’s Falcon M 2 I<br />

Falco naumanni Lesser Kestrel B 1 I<br />

Fringilla coelebs Common Chaffinch R I*<br />

Lanius collurio Red-backed Shrike M 3 I<br />

Lanius minor Lesser Grey Shrike M 2 I<br />

Lullula arborea Wood Lark R 2 I<br />

Parus ater Coal Tit R I*<br />

Pica pica Black-billed Magpie R II/2<br />

Streptopelia<br />

Eurasian Collared dove R II/2<br />

decaocto<br />

Streptopelia turtur European Turtle dove B 3 II/2<br />

Sturnus vulgaris Common Starling R 3 II/2<br />

Troglodytes<br />

Winter Wren R I*<br />

troglodytes<br />

Turdus merula Common Blackbird R II/2<br />

Turdus philomelos Song Thrush R II/2<br />

Turdus viscivorus Mistle Thrush R II/2<br />

1 B: Breeding, M: Migrant that is breeding in the area, R: Resident<br />

2<br />

SPEC 1: Species of global conservation concern.<br />

SPEC 2: Concentrated in Europe <strong>and</strong> with an Unfavourable Conservation Status.<br />

SPEC3: Not concentrated in Europe but with an Unfavourable Conservation Status.<br />

Figure 1: Boxplots for bird richness by habitat type.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.G. Plexida & A.I. Sfougaris 2010. Conservation of priority bird species in a protected area of central Greece<br />

290<br />

Figure 2: Boxplots for bird abundance by habitat type.<br />

Figure 3: Map of modeled priority bird richness produced by the ordinary kriging model <strong>and</strong> location of<br />

sampling points.<br />

St<strong>and</strong>ardized semivariance<br />

γ<br />

1,92<br />

1,54<br />

1,15<br />

0,77<br />

0,38<br />

0 0,34 0,67 1,01 1,35 1,68 2,02 2,35 2,69<br />

Distance, h 10 -4<br />

γ 10 -3 2,31<br />

1,85<br />

1,39<br />

0,93<br />

0,46<br />

0 0,43 0,85 1,28 1,7 2,13 2,56 2,98 3,41<br />

Distance, h 10 -3<br />

Figure 4: Semivariograms showing the st<strong>and</strong>arized semivariance according to the distance between<br />

surveyed points for bird (a) richness <strong>and</strong> (b) abundance in this study.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Poljanec & A. Boncina 2010. <strong>Change</strong>s in structure <strong>and</strong> composition of forest st<strong>and</strong>s at regional <strong>and</strong> national level<br />

291<br />

<strong>Change</strong>s in structure <strong>and</strong> composition of forest st<strong>and</strong>s at regional <strong>and</strong><br />

national level in the last four decades - a consequence of environmental,<br />

natural or social factors<br />

Ales Poljanec * & Andrej Boncina<br />

University of Ljubljana, Biotechnical Faculty, Department of <strong>Forest</strong>ry <strong>and</strong> Renewable<br />

<strong>Forest</strong> Resources, Vecna pot 83, 1000 Ljubljana, Slovenia<br />

Abstract<br />

Based on data acquired from the spatial information system Silva-SI, the majority of the<br />

entire forest area in Slovenia (22,220 forest compartments with a total area of 7,183<br />

km 2 ) was analysed for changes in structure (growing stock, dbh structure) <strong>and</strong> tree<br />

species composition of forest st<strong>and</strong>s in the period 1970–2008. Different statistical<br />

methods (data mining, GLM) <strong>and</strong> GIS-supported analyses were used to test influence of<br />

20 variables on changes of forest st<strong>and</strong>s: 11 environmental, 2 management, 3 st<strong>and</strong> <strong>and</strong><br />

4 social variables. In the observed period total growing stock significantly increased -<br />

from 190 to 293 m 3 /ha, as well the shares of medium-diameter (30 < dbh ≥ 50 cm) <strong>and</strong><br />

large-diameter (dbh > 50 cm) trees in the total growing stock, from 43 % to 46 % <strong>and</strong><br />

from 9 % to 18 %, respectively. Tree species composition changed significantly in the<br />

analysed period, resulting in a higher share of broadleaves, whereas the share of silver<br />

fir in total growing stock decreased from 17.5 % to 8.6 %. <strong>Change</strong>s in forest st<strong>and</strong><br />

structure were of different magnitudes <strong>and</strong> directions; their variability was partly<br />

explained by different initial states of forest st<strong>and</strong>s influenced by forest management,<br />

environmental <strong>and</strong> social factors included in the study.<br />

Keywords: forest structure, spatiotemporal dynamics, forest management, environmental<br />

factors, social factors, spatial information system<br />

1. Introduction<br />

Temperate forests in Europe cover a large bioclimatiological range <strong>and</strong> play a prominent role in<br />

timber production, nature protection, water conservation, erosion control <strong>and</strong> recreation. For<br />

centuries temperate forests in Europe have been affected by human activities. <strong>Forest</strong><br />

management has caused large-scale changes in the spatial distribution, tree species composition,<br />

<strong>and</strong> structure of forest st<strong>and</strong>s (Johann, 2007). In the 18 th <strong>and</strong> 19 th century, even-aged forestry<br />

created large areas of uniform, mainly conifer-dominated forest st<strong>and</strong>s. In the past few decades,<br />

as nature-based forestry has become widely accepted, several phenomena associated to changes<br />

in the forest structure <strong>and</strong> species distribution occurred (Spiecker, 2003; Gold et al., 2000).<br />

These phenomena are easier to recognize <strong>and</strong> explain if long term changes have been precisely<br />

documented. Archival data such as forest management plans with inventory data, forest maps,<br />

l<strong>and</strong> registers <strong>and</strong> felling records, which are often neglected source of information, enable us to<br />

quantify long-term changes <strong>and</strong> study the impacts of different factors on changes of forest<br />

structure <strong>and</strong> composition over the past decades or centuries (Chapman et al., 2006; Klopcic et<br />

* Corresponding author. Tel.:++38614231161 Fax:++3862571169<br />

Email address: ales.poljanec@bf.uni-lj.si<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Poljanec & A. Boncina 2010. <strong>Change</strong>s in structure <strong>and</strong> composition of forest st<strong>and</strong>s at regional <strong>and</strong> national level<br />

292<br />

al., 2009). The aim of this study was thus 1) to study spatiotemporal changes in forest structure<br />

<strong>and</strong> tree species composition in the last four decades at regional <strong>and</strong> national level <strong>and</strong> 2) to<br />

determinate impact of different environmental, natural <strong>and</strong> social factors on changes of forest<br />

structure <strong>and</strong> species composition.<br />

2. Methodology<br />

2.1 Study area<br />

Slovenia lies in the south-eastern part of Central Europe, between Austria, Italy, Hungary <strong>and</strong><br />

Croatia. Due to diverse environmental conditions (the Alps, the Mediterranean, the Dinaric<br />

Mountains, <strong>and</strong> the Panonian Basin), different forest management practices in the past decades<br />

<strong>and</strong> centuries <strong>and</strong> available archival data on forest st<strong>and</strong> structure <strong>and</strong> composition, Slovenia is<br />

appropriate study area to study changes in forest st<strong>and</strong> structure <strong>and</strong> composition on the<br />

l<strong>and</strong>scape level. <strong>Forest</strong>s cover 11,400 km 2 , which represents 58% of the total l<strong>and</strong> area. The<br />

underlying characteristic of the study area is a considerable variation of relief <strong>and</strong> climatic<br />

conditions (Poljanec et al., 2010). Zonality of forest vegetation in Slovenia is quite clearly<br />

defined due to distinctive orographic factors, different soil substrata <strong>and</strong> well-preserved forest<br />

structure. For the purpose of the research, forest vegetation was classified in eight forest types<br />

according to the terminology of the Ministerial Conference on the Protection of <strong>Forest</strong>s in<br />

Europe (MCPFE) (European forest types…, 2006), reflecting distinctive <strong>and</strong> unique patterns of<br />

human impacts, modification of species composition, latitudinal/altitudinal zonation of<br />

vegetation, climatic <strong>and</strong> edaphic variability, silvicultural systems applied <strong>and</strong> forest<br />

management intensity: Alpine coniferous forest (EFC 3; 225 km 2 ), acidophilous oakwood (EFC<br />

4.1; 241 km 2 ), sessile oak–hornbeam forest (EFC 5.2; 577 km 2 ), Central European<br />

submountainous beech forest (EFC 6.4; 1,901 km 2 ), Illyrian submountainous beech forest (EFC<br />

6.6; 1,505 km 2 ), Illyrian mountainous beech forest (EFC 7.4; 2,612 km 2 ), thermophilous<br />

deciduous forests (EFC 8; 77 km 2 ) <strong>and</strong> floodplain forest (EFC 12; 45 km 2 ).<br />

2.2. Data description <strong>and</strong> analysis<br />

The study is based on data acquired from the spatial information system Silva-SI (Poljanec,<br />

2008), which covers the entire forest area of Slovenia (N of compartments = 32,597;<br />

11,400 km2). The analysis included 22,220 compartments (7,183 km2) with an average area of<br />

34 ha for which reliable data on forest condition for 1970 <strong>and</strong> 2008 are available. An attributive<br />

database comprising seven dependent <strong>and</strong> 20 independent variables describing the site, forest<br />

management <strong>and</strong> social conditions of the compartments was designed. Among 20 independent<br />

variables, the first 11 describe site conditions (INC - mean incline, INC_SD - st<strong>and</strong>ard deviation<br />

of incline, ELV - mean elevation, ELV_SD - st<strong>and</strong>ard deviation of elevation, ASP - aspect,<br />

ASP_VAR - variation of aspect, ROC - proportion of area covered with rocks, BEDR - bedrock,<br />

T - mean annual temperature, PRCEP - mean annual precipitation, RK – site productivity), the<br />

next 4 are social variables (OWN - ownership, N_OWN – number of owners in the<br />

compartment, HS - holding size, SUCC - share of ab<strong>and</strong>oned l<strong>and</strong>), 3 variables describing state<br />

of the st<strong>and</strong>s in 1970 (GS1970 - growing stock in 1970, VOL_C - proportion of large-size<br />

diameter trees in growing stock, P_CON - proportion of conifers in growing stock) <strong>and</strong> the last<br />

2 variables describe the intensity of management (FMR – forest management region, CUT -<br />

annual allowable cut (m3 ha-1) in the period 1970-2008). Data were acquired from various<br />

sources, mostly through forest inventories; this is a combination of a field description of all<br />

st<strong>and</strong>s <strong>and</strong> of tree measurements (dbh ≥ 10 cm) at permanent 500 m2 sampling plots<br />

(N = 100,178) with a dominant sampling network size of 250 m × 250 m <strong>and</strong> 250 m × 500 m<br />

(Poljanec et al., 2010).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Poljanec & A. Boncina 2010. <strong>Change</strong>s in structure <strong>and</strong> composition of forest st<strong>and</strong>s at regional <strong>and</strong> national level<br />

293<br />

Differences in changes of forest st<strong>and</strong> structure <strong>and</strong> composition was investigated by eight<br />

forest types with seven variables: the variables ΔGS, ΔS, ΔM <strong>and</strong> ΔL denotes the differences in<br />

the total growing stock, growing stock of small, medium <strong>and</strong> large diameter trees between 2008<br />

<strong>and</strong> 1970, while the variables ΔP_Beech, ΔP_Fir <strong>and</strong> ΔP_Spruce denotes the differences in the<br />

proportion of European beech (Fagus sylvatica L.), silver fir (Abies alba Mill.)<strong>and</strong> Norway<br />

spruce (Picea abies (L) H. Karst.) in the growing stock of forest st<strong>and</strong>s in 1970 <strong>and</strong> 2008. The<br />

Welch test (Welch, 1947) was used due to the non-homogeneity of variances <strong>and</strong> different<br />

sample sizes <strong>and</strong> the Tamhane's T2 test (Tamhane, 1979) for post hoc multiple comparisons of<br />

mean values. The impact of selected site, st<strong>and</strong>, forest management <strong>and</strong> social factors on<br />

changes of forest st<strong>and</strong> structure <strong>and</strong> composition were investigated with multiple regression<br />

approach (GLM). All statistical analysis was carried out in the SPSS 17.0.<br />

3. Results<br />

The analyses showed that the structure of forest st<strong>and</strong>s changed significantly in the period 1970-<br />

2005. In 1970–2008 growing stock increased from 190 m 3 /ha to 293 m 3 /ha. An increase of total<br />

growing stock was registered in most of the compartments (N=19,775). In 358 compartments,<br />

however, there was no observed change in growing stock, <strong>and</strong> in 2.087 compartments growing<br />

stock dropped (Figure 1). <strong>Change</strong>s in diameter distribution are evidently reflected in the gradual<br />

ageing of st<strong>and</strong>s, as the share of small-diameter trees dropped (from 49% of total growing stock<br />

in 1970 to 31% in 2005), the share of medium-diameter (30 cm ≤ dbh < 50 cm) <strong>and</strong> largediameter<br />

trees (dbh ≥ 50 cm) increased substantially (from 43 % to 46 % <strong>and</strong> from 9 % to 18 %<br />

of total growing stock). Rather than uniform, the changes in conifers <strong>and</strong> broadleaves are<br />

characteristically different. A considerable increase in the growing stock of small- (10 cm ≤ dbh<br />

< 30 cm) <strong>and</strong> medium-diameter trees was recorded for broadleaved species, whereas for<br />

conifers an increase in the growing stock of large-diameter trees <strong>and</strong> a decrease in the growing<br />

stock of small-diameter trees is evident. Furthermore, the tree species composition of forests<br />

changed substantially in the period 1970–2005. The changes are reflected in higher share of<br />

broadleaves (rising from 40 % of total growing stock in 1970 to 48 % in 2008) <strong>and</strong> improved<br />

conservation status of forests. The growing stock of beech doubled, whereas the share of silver<br />

fir decreased from 17.5% to 8.6 % of total growing stock. <strong>Change</strong>s were less evident for spruce,<br />

resulting in smaller increase of its share in the total grooving stock (rise from 34 % to 37 % of<br />

total growing stock).<br />

<strong>Change</strong>s of growing stock were significantly different among forest types (Welch statistic =<br />

42.547, p=0.000). The increase in growing stock was the highest in EFT 6.4 <strong>and</strong> EFT 6.6, while<br />

changes of growing stock are the smallest in EFC 3 <strong>and</strong> EFC 8. The increase of growing stock is<br />

also correlated to an increase of the proportion of small- <strong>and</strong> medium-diameter trees, while the<br />

increase of large-diameter trees was the highest in forest types where changes of growing stock<br />

were the smallest (EFC 3 <strong>and</strong> EFC 7.4). Furthermore, changes of the proportion of Norway<br />

spruce (Welch statistic = 77.127, p=0.000), silver fir (Welch statistic = 182.214, p=0.000) <strong>and</strong><br />

European beech (Welch statistic = 38.780, p=0.000) were significantly different among forest<br />

types. Proportion of spruce decreased in forest types where spruce is not present in potential<br />

vegetation (EFC 12) <strong>and</strong> increased the most in mountainous vegetation belt, especially in forest<br />

types EFC 7.4 <strong>and</strong> EFC 3. The proportion of beech in the total growing stock dropped in EFC 3<br />

<strong>and</strong> EFC 8 which is particularly distinguished from the changes in other forest types where the<br />

proportion of beech increased in the observed period. The proportion of silver fir in the total<br />

growing stock dropped in all forest types. The regression process is distinctive in forest type<br />

EFC 7.4.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Poljanec & A. Boncina 2010. <strong>Change</strong>s in structure <strong>and</strong> composition of forest st<strong>and</strong>s at regional <strong>and</strong> national level<br />

294<br />

<strong>Change</strong>s of forest structure, occurred in the period 1970–2005, were mainly influenced by forest<br />

management; forest management is reflected through the harvest intensity <strong>and</strong> various forest<br />

management concepts in forest management regions. In addition, changes of forest st<strong>and</strong>s were<br />

significantly influenced by the initial state of the st<strong>and</strong>s. The increase of total growing stock was<br />

the highest in forests with lower growing stock, high proportion of broadleaves <strong>and</strong> less<br />

intensive management. In these forests the growing stock of low-diameter <strong>and</strong> mediumdiameter<br />

trees rose, too. Ownership, holding size <strong>and</strong> share of ab<strong>and</strong>oned l<strong>and</strong> in a compartment<br />

were the most important social factors, whereas the main environmental factors were altitude,<br />

mean annual temperature, mean annual precipitation <strong>and</strong> site productivity.<br />

4. Discussion<br />

<strong>Forest</strong> st<strong>and</strong>s in the study area <strong>and</strong> more broadly in the central Europe changed significantly in<br />

last centuries (Klopcic et al., 2009; Spiecker, 2003). In the observed period changes are<br />

reflected in constant increase of growing stock, general aging of forest st<strong>and</strong>s <strong>and</strong> shifting tree<br />

species composition closer to current climatic <strong>and</strong> edaphic conditions. In spite of the general<br />

trends of changing composition <strong>and</strong> structure of forest st<strong>and</strong>s, this study <strong>and</strong> several others (e.g.<br />

Poljanec et al., 2010) indicate that changes varied in nature <strong>and</strong> magnitude between different<br />

forest types <strong>and</strong> different areas. <strong>Forest</strong> resources in extreme site conditions were least affected,<br />

in particular in the mountains where forest changes are slow due to harsh growing conditions,<br />

while in other forests types, changes of forest structure <strong>and</strong> composition can take on different<br />

forms. Varying magnitude <strong>and</strong> orientation of changes in st<strong>and</strong> parameters can be linked to the<br />

impact of different site conditions (Oliver <strong>and</strong> Larson, 1996), initial state of forest st<strong>and</strong>s <strong>and</strong><br />

different natural <strong>and</strong> anthropogenic disturbances (Pickett <strong>and</strong> White, 1996), as well as the<br />

selection of indicators <strong>and</strong> scale in which changes are observed.<br />

Initial state of forest st<strong>and</strong>s in 1970 <strong>and</strong> forest management are the most important factors<br />

explaining changes in forest st<strong>and</strong> structure <strong>and</strong> composition in last four decades, while the<br />

impact of natural <strong>and</strong> social factors is indirect as they point to differing conditions for forest<br />

management. Initial state of the forest st<strong>and</strong>s in 1970 are a result of interplay between different<br />

site conditions (Oliver <strong>and</strong> Larson, 1996) <strong>and</strong> different past forest management <strong>and</strong> l<strong>and</strong> use<br />

history (Spiecker, 2003; Johann, 2007). In the most of the study area intensive harvesting <strong>and</strong><br />

even-aged forestry promoting spruce significantly change st<strong>and</strong> structure <strong>and</strong> composition,<br />

while the application of uneven-aged systems in the 19 th <strong>and</strong> beginning of 20 th century<br />

particularly in the Dinaric region resulted in more preserved forest st<strong>and</strong> structure <strong>and</strong><br />

composition. In the past few decades nature-based forestry (Diaci, 2006) was applied through<br />

intensive forest management planning on the entire study area. The harvest intensity decreased<br />

<strong>and</strong> only in exceptions reached the net annual increment, natural tree species <strong>and</strong> more diverse,<br />

site oriented forest structure was promoted through silvicultural measures.<br />

Current structure <strong>and</strong> composition of forest st<strong>and</strong>s in the study area showed significant<br />

improvement of forest st<strong>and</strong>s over the period under research. In fact, some st<strong>and</strong> parameter<br />

values such as average growing stock, current annual increment <strong>and</strong> diameter distribution came<br />

close to goal values. The development of species composition is generally favorable, since<br />

alteration of natural tree species composition decreased. Despite overall improved conservation<br />

status of forests, the status <strong>and</strong> future development of silver fir population is unfavourable<br />

(Boncina et al., 2009).<br />

In the future, regeneration <strong>and</strong> tending of appropriately structured growing stock is given<br />

priority over growing stock accumulation. On account of significant differences inside of study<br />

area a differentiated approach adapted to local forest conditions will be needed to ensure even<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Poljanec & A. Boncina 2010. <strong>Change</strong>s in structure <strong>and</strong> composition of forest st<strong>and</strong>s at regional <strong>and</strong> national level<br />

295<br />

accumulation of growing stock, its maintenance or active regeneration of forest st<strong>and</strong>s. Special<br />

focus will need to be placed on alleviating forest management risks, which depends mainly<br />

upon implementation of silvicultural, protective <strong>and</strong> other measures intended to increase<br />

resistance of forest st<strong>and</strong>s.<br />

References<br />

Boncina A., Ficko A., Klopcic M., Matijasic D., Poljanec A. 2009. Management of silver fir<br />

(Abies alba Mill.) in Slovenia. Research Reports <strong>Forest</strong>ry <strong>and</strong> Wood Science <strong>and</strong><br />

Technology, 90: 43–56.<br />

Chapman R. A., Heitzman E., Shelton M. G. 2006. Long-term changes in forest structure <strong>and</strong><br />

species composition of an upl<strong>and</strong> oak forest in Arkansas. <strong>Forest</strong> ecology <strong>and</strong><br />

management, 236: 85–92.<br />

Diaci J. 2006. Nature-based silviculture in Slovenia : origins, development <strong>and</strong> future trends.<br />

Studia forestalia Slovenica, 126: 119-131<br />

European forest types, categories <strong>and</strong> types for sustainable forest management reporting <strong>and</strong><br />

policy, 2006. European Environmental Agency Technical Report No. 9.<br />

Gold. S., Korotkov A.V., Sasse V. 2000. The development of European forest resources, 1950<br />

to 2000. <strong>Forest</strong> Policy <strong>and</strong> Economics, 8: 183–192<br />

Johann, E., 2007. Traditional forest management under the influence of science <strong>and</strong> industry:<br />

the story of the alpine cultural l<strong>and</strong>scapes. <strong>Forest</strong> Ecology <strong>and</strong> Management 249: 54–62.<br />

Klopcic M., Jerina K., Boncina A. 2009. Long-term changes of structure <strong>and</strong> tree species<br />

composition in Dinaric uneven-aged forests: are red deer an important factor European<br />

journal of forest research, DOI: 10.1007/s10342-009-0325-z.<br />

Oliver C. D., Larson B. C. 1996. <strong>Forest</strong> st<strong>and</strong> dynamics. Updated ed. New York, John Wiley &<br />

Sons: 520.<br />

Pickett S. T., White P. S. 1985. The ecology of natural disturbance of natural patch dynamics.<br />

San Diego, Academic Press: 472.<br />

Poljanec A., Ficko A., Boncina A. 2010. Spatiotemporal dynamic of European beech (Fagus<br />

sylvatica L.) in Slovenia, 1970-2005. <strong>Forest</strong> Ecology <strong>and</strong> Management, 259: 2183–2190.<br />

Spiecker, H., 2003. Silvicultural management in maintaining biodiversity <strong>and</strong> resistance of<br />

forests in Europe-temperate zone. Journal of Environmental Management, 67: 55–65.<br />

Tamhane, A.C., 1979. A comparison of procedures for multiple comparisons of means with<br />

unequal variances. Journal of the American Statistical Association, 74: 471–479.<br />

Welch, B.L., 1947. The generalization of ‘‘student’s’’ problem when several different<br />

population variances are involved. Biometrika, 34: 28–35.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Poljanec & A. Boncina 2010. <strong>Change</strong>s in structure <strong>and</strong> composition of forest st<strong>and</strong>s at regional <strong>and</strong> national level<br />

296<br />

Figure 1: <strong>Change</strong>s of growing stock (ΔGS) in period 1970-2008; light gray – increase of total growing<br />

stock, dark gray - no observed change in growing stock, black – decrease of total growing stock.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Redon & S. Luque 2010. Tengmalm owl <strong>and</strong> Pygmy owl as a surrogate for biodiversity value in the French Alps <strong>Forest</strong>s<br />

297<br />

Tengmalm owl (Aegolius funereus) <strong>and</strong> Pygmy owl (Glaussidium<br />

passerinum) as a surrogate for biodiversity value in the French Alps<br />

<strong>Forest</strong>s<br />

Mathilde Redon * & S<strong>and</strong>ra Luque<br />

CEMAGREF, Institute for Agricultural <strong>and</strong> Environmental Engineering Research, Grenoble,<br />

France<br />

Abstract<br />

The problem in biodiversity monitoring <strong>and</strong> conservation is that usually exist vast gaps in<br />

available information on the spatial distribution of biodiversity that poses a major challenge for<br />

the development of biodiversity indicators <strong>and</strong> regional conservation planning.<br />

Within this context, the concept of habitat quality is fundamental to the study of ecology.<br />

Measurements of habitat structure offer the potential to directly predict quality (in terms of<br />

physical structure <strong>and</strong> plant species composition). An example using two bird species,<br />

Tengmalm owl (Aegolius funereus) <strong>and</strong> Pygmy owl (Glaussidium passerinum) is presented as<br />

indicator of forest biodiversity. Maximum entropy (Maxent), a presence-only modelling<br />

approach, is used to model the distribution of these two species within a large study area in the<br />

French Alps. Despite biased sampling design, this method performs very well in predicting<br />

spatial distribution of the two owl species. Results are then used as criteria in the habitat<br />

biodiversity value index.<br />

Key words: Habitat quality, Maxent, distribution modelling, biodiversity indicators, French<br />

Alps<br />

1. Introduction<br />

Improving knowledge on the distribution of indicator <strong>and</strong> emblematic but locally poorly known<br />

species is of great importance for managers as well as for naturalists (Baldwin, 2009). These<br />

species can be used as a surrogate for biodiversity monitoring <strong>and</strong> conservation (Lindenmayer et<br />

al., 2000). Still vast gaps in available information on the spatial distribution of biodiversity<br />

exist, that poses a major challenge for the development of relevant biodiversity indicators for<br />

regional conservation <strong>and</strong> forest management planning.<br />

In addition, the development of spatial knowledge on the habitat requirements <strong>and</strong> ecology of<br />

these species would facilitate conservation of a great number of related species.<br />

Models that establish relationships between environmental variables <strong>and</strong> species<br />

occurrence have been developed <strong>and</strong> are widely used with many applications in conservation<br />

<strong>and</strong> management-related fields (Cowley et al., 2000; Elith et al., 2006; Gibson et al., 2004;<br />

Pearce <strong>and</strong> Ferrier, 2000; Stockwell <strong>and</strong> Peterson, 2002). They can help to guide additional field<br />

work, by identifying unknown population locations. It also supports management decisions with<br />

regard to biodiversity, to determine suitable sites for reintroductions or to assist selection of<br />

protected areas (Baldwin, 2009). First, these models were mainly developed for presenceabsence<br />

data modelling. However, absence data are often lacking or biased <strong>and</strong> a new<br />

generation of models adapted to presence-only data modelling have been proposed (Baldwin,<br />

2009)(Hirzel et al., 2002; Phillips et al., 2006). A huge number of such methods exist <strong>and</strong> differ<br />

from their data requirements, statistical models used, output formats, performance in diverse<br />

situations (Elith et al., 2006; Guisan <strong>and</strong> Zimmermann, 2000). Much of them are based on the<br />

ecological niche theory (Hirzel <strong>and</strong> Le Lay, 2008)(Phillips et al., 2006). Ecological-niche based<br />

models generally define a function that links the fitness of individuals to their environment<br />

* Corresponding author.<br />

Email address: Mathilde.redon@cemagref<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Redon & S. Luque 2010. Tengmalm owl <strong>and</strong> Pygmy owl as a surrogate for biodiversity value in the French Alps <strong>Forest</strong>s<br />

298<br />

(Hirzel <strong>and</strong> Le Lay, 2008).Thus, theoretically, if we know precisely the habitat characteristics of<br />

a species, it is possible to rebuild its ecological niche from the environmental variables<br />

describing its habitat.<br />

In this study, we have chosen to use the Maximum Entropy modelling approach, which<br />

is a relatively recent method developed by Phillips (2006). Maxent is a presence-only modelling<br />

approach with a proved good potential to predict wildlife distribution. Despite biased sampling<br />

design, this method performs very well in predicting spatial distribution of species data (Elith et<br />

al., 2006). It estimates the less constrained distribution of training points compare to r<strong>and</strong>om<br />

background locations with environmental data layers defining constrains (Baldwin, 2009). The<br />

results show how well the model fits the location data as compared to a r<strong>and</strong>om distribution<br />

(Phillips et al., 2006; Phillips et al., 2004).<br />

Herein we aim to predict the distribution of two owls’ species, Tengmalm owl (Aegolius<br />

funereus) <strong>and</strong> Pygmy owl (Glaussidium passerinum), in the French Alps. These species have<br />

specific habitat needs <strong>and</strong> their presence reflects those of numerous other forest-dwelling<br />

species. They are considered as relicts from Ice Age <strong>and</strong> need quite cold areas, which make<br />

them good c<strong>and</strong>idates to develop further studies in relation to global climate changes.<br />

In addition, their distributions are poorly known <strong>and</strong> their protection status are not well defined.<br />

It is also important to denote that Pygmy owl populations in the Vercors Mountain (Alps range)<br />

represent the occidental limit of the European range of the species. It represents also an<br />

additional stake to learn more about distribution <strong>and</strong> habitat structure of this species at the limit<br />

of its range.<br />

Requirements <strong>and</strong> distribution of Tengmalm owl are less well known because this species is<br />

nocturnal <strong>and</strong> discreet, therefore census data are difficult to gather. Modelling its potential<br />

distribution will allow to improve knowledge on its habitat <strong>and</strong> ecological needs. Several local<br />

surveys efforts took place in order to develop a census of the populations within the “Vercors”<br />

region. But these works are limited to very small areas, <strong>and</strong> a distribution model which covers<br />

the entire mountain region would be very useful to help to define adequate surveys efforts for<br />

the future while at the same time will provide an overview of the likely distribution of the two<br />

species.<br />

2. Material <strong>and</strong> methods<br />

2.1 Case study area <strong>and</strong> species occurrence data<br />

This work was conducted within Vercors Natural Regional Park (VNRP), located at the<br />

frontier between northern <strong>and</strong> southern French Alps (Figure 1). It covers 206 000 hectares with<br />

139 000 hectares of forests. Approximately a half of these forests are Public (State <strong>and</strong><br />

municipalities forests) <strong>and</strong> the rest is in the h<strong>and</strong>s of private stakeholders. The main tree species<br />

are Fir (Abies alba), Spruce (Picea abies) <strong>and</strong> Birch (Fagus sylvatica).<br />

We used Tengmalm owl (Aegolius funereus) <strong>and</strong> Pygmy owl (Glaussidium passerinum)<br />

point counts data from several surveys conducted in the ‘Hauts Plateaux du Vercors’ Natural<br />

Reserve (HPVNR), which is located within the VNRP. The Reserve is mainly composed of<br />

three main forest types: i) mixed uneven-aged birch/spruce/fir forests, ii) pure quite sparse evenaged<br />

or uneven-aged spruce forests <strong>and</strong> at high elevation, iii) pure sparse naturally even-aged<br />

Mountain Pine forests. The bigger State forest in the Reserve has recently been classified as an<br />

Integral Biological Reserve (IBR).<br />

All local surveys have take place in this State forest <strong>and</strong> mainly in the IBR. The two<br />

owls’ species are from the North European boreal forests <strong>and</strong> the cold sparse spruce Reserve<br />

forests look-like their original habitat. Therefore, local people generally thought that the range<br />

of the two species is limited to these particular forests in the Vercors.<br />

In this part of the Alps, Pygmy owl depends on cavities carved by the Great spotted woodpecker<br />

(Dendrocopus major) <strong>and</strong> Tengmalm owl by the Black Woodpecker (Dryocopus martius) for<br />

breeding.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Redon & S. Luque 2010. Tengmalm owl <strong>and</strong> Pygmy owl as a surrogate for biodiversity value in the French Alps <strong>Forest</strong>s<br />

299<br />

These cavity providers favour respectively spruce <strong>and</strong> birch trees to breed. It implies that the<br />

presence of the woodpeckers <strong>and</strong> their host trees are likely to be important habitat variables for<br />

the two owls.<br />

The point counts data come from the naturalist network of the National <strong>Forest</strong> Office<br />

<strong>and</strong> the “Ligue de Protection des Oiseaux”, an organism which aims to improve knowledge on<br />

the local fauna species.<br />

These data are a combination of visual <strong>and</strong> eared bird contacts in addition to nests locations.<br />

Each contact point is located with a <strong>Global</strong> Positioning System (GPS). The reliability of these<br />

data is very heterogeneous because each data source has its own sampling design <strong>and</strong> its own<br />

database system. We therefore harmonize data before integration into a common database.<br />

It is important to denote that despite the low precision of visual <strong>and</strong> eared occurrence data we<br />

include them into our database since these are owl’s activity centres within their territory.<br />

The resulted dataset is composed of presence points represented as latitude/longitude<br />

coordinates, then no absence points are considered. This is a common issue when someone<br />

works with wildlife surveys data (Anderson et al., 2003; Chefaoui <strong>and</strong> Lobo, 2008). Therefore<br />

the interest of models like Maxent, as aforementioned, is the use of presence data only for the<br />

computation of the habitat modelling.<br />

2.2 GIS environmental data<br />

We used a set of environmental data based on the knowledge of the species ecology <strong>and</strong><br />

factors affecting distribution of the species within the entire study area: elevation, aspect, slope,<br />

topography, forest habitats (from Alpine National Botanic Conservatory topology), related<br />

woodpeckers species presence (Dendrocopus major for Pygmy owl <strong>and</strong> Dryocopus martius for<br />

Tengmalm owl), l<strong>and</strong> cover (Corine L<strong>and</strong> Cover 2006 level 3), mean annual temperature,<br />

woodpeckers host tree species (Spruce <strong>and</strong> Birch).<br />

These data are represented as raster layers with a 50 m resolution; we used ArcGIS 9.3 to<br />

prepare the different data layers.<br />

A single raster mask delimiting the study area was used to assure that all raster layers have the<br />

same dimensions.<br />

For the two species, we used 20% r<strong>and</strong>omly selected occurrence data for cross-validation,<br />

leaving the remaining 80 % for analysis.<br />

We implemented the model with freeware Maxent developed by (Phillips et al., 2005).<br />

It is friendly use, as species occurrence training <strong>and</strong> test files <strong>and</strong> environmental data layers are<br />

automatically recognize by the application.<br />

We used simultaneously continuous <strong>and</strong> discrete data. We let almost all default<br />

parameters, but we set the regularization value to 1 for the two species. We also chose to see the<br />

jackknife test of variable importance <strong>and</strong> the response curves to evaluate the relative<br />

contribution of each variable to the model.<br />

We first include all the environmental variables in the model. Then, we delete those which did<br />

not show any significant contribution to the model.<br />

Five variables were finally selected for the two species: elevation, topography, l<strong>and</strong><br />

cover, mean annual temperature <strong>and</strong> presence of Spruce for Pygmy owl <strong>and</strong> L<strong>and</strong> cover,<br />

elevation, mean annual temperature, slope <strong>and</strong> Birch presence for Tengmalm owl.<br />

Maxent provides three output formats. We select the logistic output as generally<br />

recommended. The result is a continuous value between 0 <strong>and</strong> 100. Each resulting raster pixel<br />

contains a value reflecting how well the predictive conditions for each pixel are.<br />

We then export results into ArcGIS 9.3 in order to apply a threshold value to produce<br />

the occurrence map. Many methods exist to determine the presence threshold. We used 10<br />

percentile training presence (threshold= 0.305 for Pygmy owl <strong>and</strong> 0.227 for Tengmalm owl) as<br />

suggested by (Phillips <strong>and</strong> Dudík, 2008). This threshold value provides a better ecologically<br />

significant result when compared with more restricted thresholds values.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Redon & S. Luque 2010. Tengmalm owl <strong>and</strong> Pygmy owl as a surrogate for biodiversity value in the French Alps <strong>Forest</strong>s<br />

300<br />

3. Model evaluation<br />

To evaluate models of species distribution, the best method would have been to use an<br />

independent data set. However, for the two owl’s species, observation data are spatially<br />

aggregated <strong>and</strong> it would have had no sense to use data located in the same place than the<br />

training data to evaluate the performance of the model.<br />

We test models performance with several other tools.<br />

Maxent calculate the AUC (Area Under the receiver operating Curve) for each run. It is<br />

a st<strong>and</strong>ard, threshold-independent method for model evaluation. This method was initially<br />

developed for presence-absence data. In Maxent, absence data are replaced by r<strong>and</strong>om points<br />

(Phillips et al., 2006). AUC tests if a prediction is better than r<strong>and</strong>om for any possible presence<br />

threshold. It varies between 0.5 when the result is not better than a r<strong>and</strong>om selection <strong>and</strong> 1 when<br />

the result is significantly better than r<strong>and</strong>om.<br />

Maxent also calculates an omission rate for training <strong>and</strong> test data. Omission rate<br />

indicates the percentage of test localities that falls into pixels not predicted as suitable for the<br />

species (Phillips et al., 2006). It should be low for a good model performance.<br />

We also verify if all training points where predicted with a high probability.<br />

In addition, we compare environmental variables values between training sites <strong>and</strong> the same<br />

number of points r<strong>and</strong>omly selected among positively predicted sites. As observation data<br />

represent activity centres but rarely nesting sites, we used circles representing owls living<br />

territories to make the comparison (Hakkarainen et al., 2008; Ribe et al., 1998). Each selected<br />

point was the centre of one living territory (i.e. this include nesting sites) considered with a<br />

radius of 670 meters for Pygmy owl <strong>and</strong> of 1000 meters for Tengmalm owl. Mean of<br />

quantitative variables <strong>and</strong> dominant value of qualitative variables were used for the comparison.<br />

As data do not follow a Gaussian distribution, we used non-parametric statistic tests<br />

implemented in R 2.8.1 (Gentleman & Ihaka, 1997).<br />

4. Results <strong>and</strong> discussion<br />

Despite biased sampling design, Maxent model performed very well in predicting<br />

potential spatial distribution of the two owl species. Test omission rates are null at minimum<br />

training presence threshold (0.000 for Pygmy owl <strong>and</strong> Tengmalm owl) <strong>and</strong> low at 10 percentile<br />

training presence threshold (0.176 For Pygmy owl <strong>and</strong> 0.067 for Tengmalm owl).<br />

For the two species, models show an AUC value very close to 1. However, when species have a<br />

narrow range (or training data are spatially aggregated), AUC is overestimated (Phillips <strong>and</strong><br />

Dudík, 2008), which is certainly the case here.<br />

Mean training data predictive rate is 0.58 (SD= 0.20) for Pygmy owl <strong>and</strong> 0.55 (SD= 0.20) for<br />

Tengmalm owl. The model performance seems to be good then based on the well predicted<br />

calibration points. In addition, for the two species, there are no significant differences in<br />

environmental variables values between training <strong>and</strong> r<strong>and</strong>omly selected species living<br />

territories. It indicates that habitat is suitable within areas where the species occurrences are<br />

predicted.<br />

For the two species, the resulting distribution is wider than would be expected by local<br />

knowledge (see Figure 2). This result is not surprising because some observations have been<br />

done in Vercors forests outside of the HPVNR <strong>and</strong> in an adjacent mountain massif with a<br />

different kind of forest habitats (i.e. more humid, productive <strong>and</strong> with closed canopy<br />

conditions).<br />

The distribution maps bring new information on these poorly known species. They<br />

represent very useful data on owl species probability presence with a grain that allow their use<br />

at different scales.<br />

However, it is important to note that species could not be present in a site even if they<br />

are predicted. Other factors, not taken into account in the analysis, can explain species absence.<br />

They can be for example: predator presence (notably Tawny owl (Strix aluco) for the two owls<br />

<strong>and</strong> European pine marten (Martes martes) for Tengmalm owl), sites far from existing<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Redon & S. Luque 2010. Tengmalm owl <strong>and</strong> Pygmy owl as a surrogate for biodiversity value in the French Alps <strong>Forest</strong>s<br />

301<br />

population <strong>and</strong> not yet colonized <strong>and</strong> a lack of prey resources (little mammals, passerine birds,<br />

etc.).<br />

Therefore, sites of predicted presence would guide naturalists’ future work in order to<br />

identify other suitable sites were the bird distribution is unknown while at the same time<br />

facilitate selection of areas with high ecological value. As numerous public forests are managed<br />

for wood production in the Vercors, these maps would allow to better integrate biodiversity<br />

conservation into management planning. In addition, as these species are linked to cold habitats,<br />

they could serve as good indicators of climate change with further work including temporal<br />

analysis. The kind of models used here would be useful to follow the evolution of their spatial<br />

distribution in years to come. Furthermore, the tools developed can be applied in assessing<br />

biodiversity value of both managed <strong>and</strong> protected forest areas to help decision-making<br />

concerning the protection of valuable habitats. Distribution modelling of these species is among<br />

the first attempts to model suitable habitat distribution of cavity-nesting owl species in France.<br />

We hope it will launch the use of such methods, which aim to improve species ecological<br />

knowledge <strong>and</strong> facilitate species censuses <strong>and</strong> conservation.<br />

Figures<br />

Figure 1: Study area localisation<br />

Figure 2: Result of Maxent model for Pygmy owl, with 10 percentile training presence threshold<br />

(left) <strong>and</strong> Tengmalm owl, with 10 percentile training presence threshold (right). Darker tones<br />

indicate higher probability of occurrence.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Redon & S. Luque 2010. Tengmalm owl <strong>and</strong> Pygmy owl as a surrogate for biodiversity value in the French Alps <strong>Forest</strong>s<br />

302<br />

References<br />

Anderson, R.P., Lew, D. <strong>and</strong> Peterson, A.T., 2003. Evaluating predictive models of species'<br />

distributions: criteria for selecting optimal models. Ecological Modelling, 162(3): 211-<br />

232.<br />

Baldwin, R.A., 2009. Use of Maximum Entropy Modeling in Wildlife Research. Entropy, 11:<br />

854-866.<br />

Chefaoui, R.M. <strong>and</strong> Lobo, J.M., 2008. Assessing the effects of pseudo-absences on predictive<br />

distribution model performance. Ecological Modelling, 210: 478-486.<br />

Cowley, M.J.R., Wilson, R.J., Leon-Cortés, J.L., Gutiérrez, D., Bulman, C.R. <strong>and</strong> Thomas,<br />

C.D., 2000. Habitat-based statistical models for predicting the spatial distribution of<br />

butterflies <strong>and</strong> day-flying moths in a fragmented l<strong>and</strong>scape. Journal of Applied Ecology,<br />

37: 60-72.<br />

Elith, J., Graham, C.H., Anderson, R.P., Dudík, M., Ferrier, S., Guisan, A., Hijmans, R.J.,<br />

Huettmann, F., Leathwick, J.R., Lehmann, A., Li, J., Lohmann, L.G., Loiselle, B.A.,<br />

Manion, G., Moritz, C., Nakamura, M., Nakazawa, Y., Mc Overton, C.O., Peterson,<br />

A.T., Phillips, S.J., Richardson, K., Scachetti-Pereira, R., Schapire, R.E., Soberón, J.,<br />

Williams, S., Wisz, M.S. <strong>and</strong> Zimmerman, N.E., 2006. Novel methods improve<br />

prediction of species’ distributions from occurrence data. Ecography, 29: 129-151.<br />

Gibson, L.A., Wilson, B.A., Cahill, D.M. <strong>and</strong> Hill, J., 2004. Spatial prediction of rufous<br />

bristlebird habitat in a coastal heathl<strong>and</strong>: a GIS-based approach. Journal of Applied<br />

Ecology, 41: 213-223.<br />

Guisan, A. <strong>and</strong> Zimmermann, N.E., 2000. Predictive habitat distribution models in ecology.<br />

Ecological Modelling, 135(2-3): 147-186.<br />

Hakkarainen, H., Korpimäki, E., Laaksonen, T., Nikula, A. <strong>and</strong> Suorsa, P., 2008. Survival of<br />

male Tengmalm’s owls increases with cover of old forest in their territory. Oecologia,<br />

155: 479-486.<br />

Hirzel, A.H., Hausser, J., Chessel, D. <strong>and</strong> Perrin, N., 2002. Ecological-niche Factor Analysis :<br />

How to compute habitat-suitability maps without absence data Ecology, 83(7): 2027-<br />

2036.<br />

Hirzel, A.H. <strong>and</strong> Le Lay, G., 2008. Habitat suitability modelling <strong>and</strong> niche theory. Journal of<br />

Applied Ecology, 45: 1372-1381.<br />

Lindenmayer, D.B., Margules, C.R. <strong>and</strong> Botkin, D.B., 2000. Indicators of Biodiversity for<br />

Ecologically Sustainable <strong>Forest</strong> Management. Conservation Biology, 14(4): 941-950.<br />

Pearce, J. <strong>and</strong> Ferrier, S., 2000. An evaluation of alternative algorithms for fitting species<br />

distribution models using logistic regression. Ecological Modelling, 128(2-3): 127-147.<br />

Phillips, S.J., Anderson, R.P. <strong>and</strong> Schapire, R.E., 2006. Maximum entropy modeling of species<br />

geographic distributions. Ecological Modelling, 190: 231-259.<br />

Phillips, S.J., Dud´ık, M. <strong>and</strong> Schapire, R.E., 2004. A Maximum Entropy Approach to Species<br />

Distribution Modeling, Proceedings of the 21st International Conference on Machine<br />

Learning, Banff, Canada.<br />

Phillips, S.J. <strong>and</strong> Dudík, M., 2008. Modeling of species distributions with Maxent: new<br />

extensions <strong>and</strong> a comprehensive evaluation. Ecography, 31: 161-175.<br />

Ribe, R., Morganti, R., Hulse, D. <strong>and</strong> Shull, R., 1998. A management driven investigation of<br />

l<strong>and</strong>scape patterns of northern spotted owl nesting territories in the high Cascades of<br />

Oregon. L<strong>and</strong>scape Ecology, 13: 1-13.<br />

Stockwell, D.R.B. <strong>and</strong> Peterson, A.T., 2002. Effects of sample size on accuracy of species<br />

distribution models. Ecological Modelling, 148(1): 1-13.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C.C. Rodrigues et al. 2010. Contribution to the characterization of Gentiana pneumonanthe L. <strong>and</strong> Maculinea alcon L.<br />

303<br />

Contribution to the characterization of Gentiana pneumonanthe L.<br />

<strong>and</strong> Maculinea alcon L. distribution in the Alvão Natural Park<br />

M. da Conceição C. Rodrigues 1,2 , Paula M. S. O. Arnaldo 1,2 & José Aranha 1,2*<br />

1<br />

Departamento de Ciências Florestais e Arquitectura Paisagista, Universidade de Trásos-Montes<br />

e Alto Douro, 5001-801 Vila Real, Portugal<br />

2<br />

CITAB, Universidade de Trás-os-Montes e Alto Douro, 5001-801 Vila Real, Portugal<br />

Abstract<br />

The aim of this research was to study the distribution of Gentiana pneumonanthe in the<br />

geographic area of Alvão Natural Park, in order to determine the area of Maculinea alcon L.<br />

expansion as well the region’s potential for development <strong>and</strong> conservation this butterfly species.<br />

Due to the ecology needs of Maculinea alcon L (simultaneous presence of G. pneumonanthe -<br />

the host plant - <strong>and</strong> Myrmica sp. - Ant that feeds the larvae of M. alcon in its larvae stadium),<br />

we made a survey of plants, butterflies <strong>and</strong> ants’ nests, using a DGPS. Data collected during<br />

filed work was then used to create a GIS, in order to analyse the relationship between plants,<br />

ants <strong>and</strong> butterflies.<br />

The results show that the l<strong>and</strong> with less human activity are the most favourable to the plants<br />

development as well to the ants development <strong>and</strong>, therefore, to the presence of butterflies.<br />

Keywords: Maculinea alcon L, Gentiana pneumonanthe, Myrmica sp, Alvão, GIS<br />

1. Introduction<br />

The study <strong>and</strong> monitoring of Maculinea alcon (Dennis & Schiffermüller 1775) meta-population<br />

of, in Parque Natural do Alvão – Northern Portugal - began in 2006. This is a butterfly of the<br />

family Lycaenidae Lepidoptera, with a life cycle very sui geniris <strong>and</strong> which is listed in the Red<br />

Book of endangered species (Swaay et al 1999). The host of the early life of the larva, of this<br />

butterfly, is a plant, Gentiana pneumonanthe L., <strong>and</strong> its host in the final larval stage is an ant in<br />

the order Hymenoptera <strong>and</strong> the family Formicidae, which is genus <strong>and</strong> species is Myrmica<br />

Alobar (Forel 1909 ).<br />

According Nowicki et al. 2005, based on studies developed in southern Pol<strong>and</strong>, in wet meadows<br />

in the valley of the Vistula, egg-laying by females of Maculinea alcon, has a high correlation<br />

with the number of flowers that each plant Gentiana pneumonanthe has during flowering season.<br />

Thomas <strong>and</strong> Elmes 2001, in the Valley Tidna, Cornwall, confirmed that each species of<br />

Maculinea has a remarkably short period of oviposition in relation to the growth stage of<br />

Gentiana pneumonanthe. According the research results, these researchers state that Maculinea<br />

teleius <strong>and</strong> Maculinea nausithous have different sites of oviposition, which is also related with<br />

the local distribution of ant host species. These two species of Maculinea have preference for<br />

floral buds, to deposit their eggs in places where it occurs Myrmica scabrinodis with a short<br />

vegetation cover between 0 <strong>and</strong> 30 cm in the case of Maculinea teleius while Maculinea<br />

nausithous prefers the flower buds with neighborhood predominant Myrmica rubra <strong>and</strong> plant<br />

strata higher.<br />

* Corresponding author; Telf. + 351 259 350 856 - Fax. + 351 250 350 480<br />

Email address: j.aranha.utad@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C.C. Rodrigues et al. 2010. Contribution to the characterization of Gentiana pneumonanthe L. <strong>and</strong> Maculinea alcon L.<br />

304<br />

In studies developed in central Europe by Schlick-Steiner et al., 2004, the author concluded that<br />

Maculinea rebellion, produces or secretes a complex substance whose composition is based on<br />

carbohydrates, which attracts the ants of the genus Myrmica. These ants, when found the larvae,<br />

in their third instar, in soil, take <strong>and</strong> lead them to the ants’ nest <strong>and</strong> feed them over the winter<br />

until the pupal stage <strong>and</strong> chrysalis. Over this instar, there is no segregation of that substance <strong>and</strong><br />

Maculinea rebellion butterfly must leave ants’ nest <strong>and</strong> fly out.<br />

This secretion, has similar characteristics to the ants odoriferous, which allows the larvae to be<br />

treated as pairs, inside the ants’ nest <strong>and</strong> is thus safeguarded from serving food to the colony. In<br />

that case, the substance studied, shows a capacity citron multiple of mimicry to the odorous of<br />

the Myrmica species of the region (M. sabuleti <strong>and</strong> M. schenck), presenting himself as a<br />

chemical strategy to guarantee that at least one of those species, take responsibility for their<br />

introduction.<br />

The aim of this work is to create a GIS to Parque Natural do Alvão, for mapping the current<br />

geographical distribution of Gentiana pneumonanthe L.<strong>and</strong> for mapping the location of<br />

Maculinea alcon meta-populations, as well traces of their presence (e.g. eggs on flowers)<br />

2. Methodology<br />

In October 2008 <strong>and</strong> March 2009, research team performed an intensive field work for data<br />

collection <strong>and</strong> survey, in order to updating GIS with the distribution of Gentiana pneumonanthe<br />

L., the location of nests of Myrmica alobar, the location of Maculinea alcon eggs on flowers<br />

<strong>and</strong> the capture of butterflies.<br />

This field work was supported by a GIS (Geographical Information System) installed on a PDA<br />

(Personal Digital Assistant) with DGPS (Differential <strong>Global</strong> Positioning System).<br />

3. Result<br />

Gentiana pneumonanthe grows well in shrub l<strong>and</strong>, with good exposure to sunlight, gentle slopes<br />

<strong>and</strong> proximity of running water. In the fringe of agricultural l<strong>and</strong>, close to irrigation channels,<br />

which minimises the effects of frost in the winter, it was also noticed its presence.<br />

In almost all the observation, it was notice the presence of Quercus robur <strong>and</strong> Quercus<br />

pyrenaica trees in the surrounding area.<br />

Many mapped Gentiana pneumonanthe plants evidenced Maculinea alcon eggs presence or<br />

eaten flowers.<br />

It was also noticed that Gentiana pneumonanthe grows well in ab<strong>and</strong>oned agricultural l<strong>and</strong>, but<br />

the development of high shrubs (Genista sp., Erica sp.) make difficult the flowers’ access by<br />

Maculinea alcon, which was evidenced by the eggs’ absence.<br />

4. Discussion<br />

Results from field work <strong>and</strong> GIS data management <strong>and</strong> processing enable to state that<br />

Maculinea alcon ecology is characterised by shrub l<strong>and</strong>, with good exposure to sunlight, gentle<br />

slopes <strong>and</strong> proximity of running water, with Quercus robur <strong>and</strong> Quercus pyrenaica trees in the<br />

surrounding area.<br />

Due to Maculinea alcon butterfly dimension <strong>and</strong> flying characteristics, high shrubs make<br />

difficult to access flowers <strong>and</strong>, this way, eggs’ deposition on flowers.<br />

References<br />

Árnyas, E., Bereczki, J., Tóth, A., Varga, Z., 2005. Results of the mark-release-recapture<br />

studies of a Maculinea rebeli population in the Aggtelek karst (N Hungary) between 2002-<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C.C. Rodrigues et al. 2010. Contribution to the characterization of Gentiana pneumonanthe L. <strong>and</strong> Maculinea alcon L.<br />

305<br />

2004. In: Studies on the Ecology <strong>and</strong> Conservation of Butterflies in Europe. Vol2: Species<br />

ecology along a European gradient: buterflies as a model. Edited by J. Settele, E. Kuhn<br />

ans J. Thomas: 111-114.<br />

Birgit C. Schlick-Steiner; Florian M. Steiner; Helmut Höttinger; Alexej Nikirov; Robert<br />

Mistrik; Christa Schafellner; Peter Baier; Erhard Christian. 2004. A butterfly’s<br />

chemical key to various ant forts: intersection-odour or aggregate-odour multihost<br />

mimicry In: Naturwissenshchaften.<br />

Irma Wynhoff, 1998: At home on foreign meadows. The reintroduction of two<br />

Maculinea butterfly species. 236 pags. Wageningen University, Bornsesteeg 69,<br />

6708 PD Wageningen, The Netherl<strong>and</strong>s. (ISBN 90-5008-406-2).<br />

J. A. Thomas <strong>and</strong> G. W. Elmes. Food-plant niche selection rather than the presence of<br />

ant nests explains ovipotion patterns in the myrmecophilous butterfly genus<br />

Maculinea. 2001. In: The Royal Society.<br />

J. Gerard B. Oostermeijer, Sheila H. Luijtem, Zdenka V. Křenová, And Hans C. M. Den<br />

Njis. 1998. Relationships between Population <strong>and</strong> Habitat Characteristics <strong>and</strong><br />

Reproduction of the Rare Gentiana pneumonanthe L. Conservation Biology. Pags<br />

1042 – 1053 Volume 12, Nº 5, October<br />

Munguira, M.L. <strong>and</strong> Martin, J., 1999. Action Plan for Maculinea Butterflies in Europe. In:<br />

Council of Europe Publishing, editor. Convention on the Conservation of European<br />

Wildlife <strong>and</strong> Natural Habitats (Bern Convention), Nature <strong>and</strong> Environment, 97, Strasburg.<br />

Piotr Nowcki, Aleks<strong>and</strong>ra Pepkowska, Joanna Kudlek, Piotr Skórka, Magdalena Witek,<br />

Josef Settele, Michal Woyciechowski. 2007. From metapopulation theory to<br />

conservation recommendations: Lessons from Spatial occurrence <strong>and</strong> abundance<br />

patterns of Maculinea butterflies. Biological Conservation. 140. Pags. 119-129.<br />

Piotr Nowicki, Magdalena Witek, Piotr Skórka, Michal Woyciechowski. 2005.<br />

Oviposition patterns in the myrmecophilous butterfly Maculinea alcon Dennis &<br />

Schiffermüller (Lepidoptera: Lycaenidae) in relation to characteristics of<br />

foodplants <strong>and</strong> presence of ant hosts. In: Polish Journal of Ecology.<br />

Thomas, J.A., 1995. The ecology <strong>and</strong> conservation of Maculinea arion <strong>and</strong> other European<br />

species of large blue butterfly. In: A.S. Pullin, editor. Ecology <strong>and</strong> Conservation of<br />

Butterflies. London: Chapman & Hall. p. 180-197.<br />

Van Swaay, C.A.M. <strong>and</strong> Warren, M.S., 1999. Red Data Book of European Butterflies<br />

(Rhophalocera). Nature <strong>and</strong> Environment 99. Council of Europe Publishing, Strasbourg<br />

Wynhoff, I. 1998. The recent distribution of the European Maculinea species. J. Insect<br />

Conservation, 2: 15-27.<br />

Acknowledgement<br />

Authors woul like to expresse is acknowledge to Fundação para a Ciência e Tecnologia (FCT),<br />

project REEQ/1163/AGR/2005 <strong>and</strong> CITAB - UTAD who support this work.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Silva et al. 2010. Using Business <strong>and</strong> Biodiversity to put Conservation into practice<br />

306<br />

Using Business <strong>and</strong> Biodiversity to put Conservation into practice: The<br />

Herdade do Esporão Case study<br />

M.C. Silva * , S. Antunes, N. Oliveira & F. Gouveia<br />

AmBioDiv – Valor Natural, Rua Filipe da Mata, Nº 10-1ºF, 1600-071 Lisbon, Portugal<br />

Abstract<br />

Although Esporão estate has a great potential to include High Conservation Value areas, this<br />

will only happen if adequate Biodiversity management <strong>and</strong> ecological restoration are put into<br />

practice. The results will be presented in the form of indicators relating l<strong>and</strong> use <strong>and</strong> the type of<br />

l<strong>and</strong>scape. The l<strong>and</strong>scape is dominated by vineyards, olive groves, oakl<strong>and</strong>s, streamside<br />

galleries, scrubl<strong>and</strong>s <strong>and</strong> grassl<strong>and</strong>s. The vegetation units of Esporão are assessed using<br />

phytosociology (Braun-Blanquet, 1979) <strong>and</strong> cartography, which are the bases for the ‘Habitat<br />

Approach’ methodlogy. The forested l<strong>and</strong>scape is quite diverse, consisting mainly in 750 ha of<br />

Holm oak Montado (Quercus rotundifolia Lam.) although most areas are extremely degraded as<br />

a result of inadequate stone pine afforestation within the Montado areas. Also highly relevant is<br />

the streamside gallery <strong>and</strong> mixed woodl<strong>and</strong> patchwork covering the area of the Caridade stream<br />

<strong>and</strong> subsidiaries, which include several High Conservation Value areas spread across the whole<br />

estate.<br />

Keywords: Biodiversity management, ecological restoration, High Conservation Value Areas,<br />

phytosociology, Habitat Approach<br />

1. Introduction<br />

The Esporão estate has signed the Business & Biodiversity (B&B) protocol in 2007, in order to<br />

promote Biodiversity <strong>and</strong> with the compromise that their activities wouldn’t affect the Esporão<br />

natural values. As a br<strong>and</strong>, Esporão is dedicated to producing premium wine <strong>and</strong> olive oil <strong>and</strong><br />

is situated (Figure 1), according to Costa et. al. (1998), between the biogeographic superdistricts<br />

Baixo Alentejano <strong>and</strong> Alto Alentejano (Mariânico-Monchiquense Sector). As a result of a fouryear<br />

field-work surveys of natural <strong>and</strong> agricultural units of l<strong>and</strong>scape, which have resulted in a<br />

formulated Biodiversity Action Plan (BAP), we will present <strong>and</strong> describe Esporão natural<br />

heritage <strong>and</strong> the main goals of the Project. The Esporão BAP phase I was finished in November<br />

2008, <strong>and</strong> is by now in the second year of monitoring. A BAP is a management tool that a)<br />

evaluates <strong>and</strong> monitors wildlife <strong>and</strong> habitats with regional/local interest, with conservation<br />

status (IUCN/ICN Red Book) <strong>and</strong> included in EU Directives, b) evaluates species with<br />

importance in crop protection <strong>and</strong> soil conservation; c) defines biological indicator groups to<br />

assess <strong>and</strong> monitor the performance of pro-Conservation practices <strong>and</strong> c) target both crop areas<br />

<strong>and</strong> surroundings, including woodl<strong>and</strong>s, wetl<strong>and</strong>s set-aside areas, inter alia, for proper habitat<br />

management. A BAP focus strongly in the concept of High Conservation Value Areas (HCVA).<br />

HCVA are l<strong>and</strong>scape level units with important natural values, i.e., habitats, fauna, flora, <strong>and</strong><br />

frequently occur in agroforestry scenarios.<br />

* Corresponding author: Tel.: 914269903<br />

Email address: mctavares@ambiodiv.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Silva et al. 2010. Using Business <strong>and</strong> Biodiversity to put Conservation into practice<br />

307<br />

2. Methodology<br />

In spite of the estate dimension, we needed to develop a useful <strong>and</strong> scientific approach which<br />

was successful to our work. So, our main methodology, designated by ‘Habitat Approach’, is<br />

focused on assessing the structure <strong>and</strong> development of an habitat mainly through the analysis of<br />

plant communities (Braun-Blanquet, 1979). A phytosociological approach can integrate the<br />

environmental variables <strong>and</strong> summarize practically the whole floristic diversity as well as many<br />

of the ecological relationships between the different organisms (Loidi, 1994). Further on, we’ve<br />

got to know if the habitat structure provides area, shelter, food <strong>and</strong>/or mutualism areas for<br />

different fauna species. At the end, we analyze the whole Esporão estate at a macro scale level,<br />

through the units of l<strong>and</strong> use <strong>and</strong> conservation types, to see if these areas act as ecological<br />

corridors. GIS was used for mapping all the habitats, plant communities, RELAPE (rare,<br />

endemic, localized, threat <strong>and</strong> endangered) species. According to Lengyel et. al (2008), habitat<br />

monitoring scheme should be based on remote sensing to record changes in l<strong>and</strong> cover <strong>and</strong><br />

habitat types, with complementary field mapping.<br />

Figure 1: Habitat approach scheme<br />

3. Results<br />

The Biodiversity Action Plan (BAP) includes the flora, habitats <strong>and</strong> fauna assessment, <strong>and</strong> also<br />

the evaluation of the vineyard’s natural enemies, as well as definition of High Conservation<br />

Value Areas (HCVA) <strong>and</strong> agroforestry management. There were identified 11 RELAPE species,<br />

namely: 6 orchids which are listed in CITES (Serapias strictiflora Welw., Serapias lingua L.,<br />

Serapias parvifloraParl., Orchis papilionacea L., Orchis morio L. <strong>and</strong> Ophrys tenthredinifera<br />

Willd.), 1 Directive Habitat specie (Narcissus bulbocodium L.), 3 european endemism (Linaria<br />

spartea (L.) Chaz, Verbascum thapsus L., Phlomis lychnitis L.) <strong>and</strong> 2 iberian endemisms<br />

(Genista polyanthos R. Roem., Narcissus jonquilla L.).<br />

We will describe the main l<strong>and</strong> uses <strong>and</strong> the Habitats Directive 92/43/CEE of Esporão Estate<br />

(Table 1), which can contain different varieties of vegetation types with different conservation<br />

values, as seen bellow. The current paper will focus only on plant ecology, leaving the fauna<br />

analysis for a next paper.<br />

Vineyards <strong>and</strong> Olive groves<br />

As wine <strong>and</strong> olive oil production is the main business of Esporão, it was necessary to make an<br />

evaluation of these areas. A lack of proper ecological structures was found, namely habitats for<br />

enhancing natural enemies. Some of these structures are the Beetle Banks <strong>and</strong> usually are made<br />

of flowers, from gramineae <strong>and</strong> compositae family, <strong>and</strong> are used mainly as a form of biological<br />

pest control to reduce the use of pesticides <strong>and</strong> insecticides. So, restoring grassl<strong>and</strong> margins in<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Silva et al. 2010. Using Business <strong>and</strong> Biodiversity to put Conservation into practice<br />

308<br />

the Esporão Vineyards <strong>and</strong> Olive groves is one of the main goals to achieve biodiversity in this<br />

l<strong>and</strong>scape unit, through ecological restoration of the set-asides areas <strong>and</strong> by installing meadow<br />

<strong>and</strong> grassl<strong>and</strong> patches. In order to reduce tillage, we also propose the use of grazers, in this case<br />

sheep, to control the weeds.<br />

Table 2: Types of l<strong>and</strong> use in Esporão Estate<br />

L<strong>and</strong> use class<br />

Total Area (ha)<br />

Council Directive 92/43/CEE habitat code<br />

Vineyards<br />

450 N/A<br />

Olive groves<br />

95 N/A<br />

Holm oak Montado (Parkl<strong>and</strong>s)<br />

750 6310 – Montados with evergreen Quercus spp.<br />

Holm oak Woodl<strong>and</strong>s 28<br />

9340 – Quercus ilex <strong>and</strong> Quercus rotundifolia<br />

forests<br />

Stone Pine plantations<br />

128 N/A<br />

Stone Pine plantations with sparse<br />

Holm oak<br />

87 N/A<br />

Orchid meadows 2.68<br />

6210 – Semi-naturaldry grassl<strong>and</strong>s <strong>and</strong> scrubl<strong>and</strong><br />

facies on calcareous substrates (Festuco-<br />

Brometalia) (* important orchid sites)<br />

Grassl<strong>and</strong>s<br />

83 N/A<br />

8230 – Siliceous rock with pioneer vegetation of<br />

the Sedo-Scleranthion or of the Sedo albi-<br />

Veronicion dillenii<br />

3270 – Rivers with muddy banks with<br />

Chenopodion rubri p.p. <strong>and</strong> Bidention p.p.<br />

3150 – Natural eutrophic lakes with<br />

Magnopotamion or Hydrocharition - type<br />

vegetation<br />

3140 – Hard oligo-mesotrophic waters with benthic<br />

vegetation of Chara spp.<br />

92D0 – Southern riparian galleries <strong>and</strong> thickets<br />

(Nerio-Tamaricetea <strong>and</strong> Securinegion tinctoriae)<br />

91B0 – Thermophilous Fraxinus angustifolia<br />

woods<br />

Streamside woodl<strong>and</strong>s <strong>and</strong> grassl<strong>and</strong>s 189<br />

3260 – Water courses of plain to montane levels<br />

with the Ranunculion fluitantis <strong>and</strong> Callitricho-<br />

Batrachion vegetation<br />

Rosemary scrubl<strong>and</strong><br />

0.16 N/A<br />

Gardens<br />

5 N/A<br />

Social areas<br />

145 N/A<br />

Holm Oak Montado (Parkl<strong>and</strong>s) <strong>and</strong> Holm Oak Woodl<strong>and</strong>s<br />

The Holm Oak Montado, a parkl<strong>and</strong> like system, represents the largest l<strong>and</strong>scape unit of<br />

Esporão estate. It consists in similar distinct types of l<strong>and</strong> use which promote annuals <strong>and</strong><br />

perennials grassl<strong>and</strong>s <strong>and</strong> scrubl<strong>and</strong>s. In these areas grazing was redefined in order to preserve<br />

sensitive areas like closed woods <strong>and</strong> wetl<strong>and</strong>s. Some of these areas were converted through<br />

stone pine afflorestation as a result of an inadequate policy actually promoted by the Portuguese<br />

governments. In historical times, an edapho-climatic climax forest must have existed in Esporão<br />

estate, but almost all these climax habitats have disappeared due to convertion, mismanagement<br />

<strong>and</strong> fires. Although all the human pressure, a few habitat patches remain defined by Pyro<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Silva et al. 2010. Using Business <strong>and</strong> Biodiversity to put Conservation into practice<br />

309<br />

bourgaeanae-Quercetum rotundifoliae woodl<strong>and</strong>s. These few areas are restricted to<br />

streamsides <strong>and</strong> damp areas which display enormous diversity. Pyro bourgaeanae-Quercetum<br />

rotundifoliae, is a s<strong>and</strong>y meso-Mediterranean, dry to sub-humid Holm oakl<strong>and</strong> which occurs in<br />

the Luso-Extremadurense Province (Pereira, 2004). Areas were this plant community occurs<br />

have a good conservation status despite their size, <strong>and</strong> are characterized by the presence of<br />

Quercus rotundifolia Lam. <strong>and</strong> Pyrus bourgaeana Decne.. Currently, there is some significant<br />

effort to restore these areas , being the main goal providing food <strong>and</strong> shelter for fauna<br />

populations, by planting myrtle, rosemary, lavender, edible fig, hawthorn, strawberry tree <strong>and</strong><br />

elmleaf blackberry species.<br />

Synthetic table of the Pyro bourgaeanae-Quercetum rotundifoliae Rivas-Martínez 1987, at<br />

slope of Caridade streamside, slopes with high inclination, N, 100 m². Characteristics: Quercus<br />

rotundifolia 2, Pyrus bourgeana 1, Olea sylvestris 1, Thapsia villosa; Companions: Cytisus<br />

scoparius 2, Cistus ladanifer 2, Lav<strong>and</strong>ula luisieri 1, Asphodelus ramosus 1, Umbilicus<br />

rupestris +, Allium massaessylum +.<br />

Stone Pine Plantations <strong>and</strong> Stone Pine Plantation with sparse Holm oak<br />

As referred above, to the government decision of financing afforestation projects was allegedly<br />

to get more l<strong>and</strong> productivity <strong>and</strong> environmental gains (Biodiversity, carbon, soil…). Due to<br />

heavy soil preparation, excessive tillage <strong>and</strong> excessive tree density the Holm oak trees present in<br />

these areas are dying, suggesting that their roots were damaged, destroyed <strong>and</strong> infected. We<br />

suggest progressive removal of all Stone Pine, good management of scrubl<strong>and</strong> <strong>and</strong> investigating<br />

Holm oak decline for further reforestation.<br />

Orchid meadows <strong>and</strong> grassl<strong>and</strong>s<br />

In the Esporão estate the orchids are present in open grassl<strong>and</strong>s <strong>and</strong> meadows between February<br />

<strong>and</strong> June. These are managed for low grazing intensity, since they are mainly present in<br />

sensitive areas. This allows orchids to grow <strong>and</strong> seed to provide a pretty ‘postcard’ for the<br />

visitors as well as raises the attention of botanists. In these habitats six orchid species can be<br />

found (Serapias strictiflora, Serapias lingua, Serapias parviflora, Orchis papilionacea, Orchis<br />

morio <strong>and</strong> Ophrys tenthredinifera). Orchid genus Serapias <strong>and</strong> Ophrys metapopulation reached<br />

over 20 individuals, making this a HCVA. This is a hemicriptofit, meso-xerofitic meadow that<br />

grows in shallow soils rich in bases <strong>and</strong> characterized by the presence of Phlomido lychnitidis-<br />

Brachypodietum phoenicoidis communities. At this point, it is important to reinforce the<br />

importance of this habitat for conservation, being a priority (as listed in the Council Directive<br />

92/43/EEC) if one of the following criteria is observed: rich orchid composition (> 4 species);<br />

presence of an important population (> 20 individuals) of one or more orchid species.<br />

Synthetic table of the Phlomido lychnitidis­Brachypodietum phoenicoidis Br.‐Bl., P. Silva &<br />

Rozeira 1956, at edges of tracks, with low inclination, E, 10 m². Characteristics: Phlomis<br />

lychnitis 4; Companions: Paronychia argentea 2, Gyn<strong>and</strong>riris sysrinchium 1, Vulpia ciliata 1.<br />

Streamside woodl<strong>and</strong>s <strong>and</strong> grassl<strong>and</strong>s<br />

This is possibly the clearest example of the ‘Habitat approach’ importance to our work. The<br />

habitat evaluation shows the right conditions for fauna settlement. Despite this first analysis, the<br />

amphibian populations that should be present in the Caridade stream were not settled (to be<br />

further explored in upcoming paper). The main reasons are suspected to be high populations of<br />

red swamp crayfish (Procambarus clarkii) <strong>and</strong> low input of water from the dam discharge. Red<br />

swamp crayfish, an alien species, prefers marshes, ponds <strong>and</strong> slow moving rivers <strong>and</strong> streams.<br />

They are tolerant to fluctuating water levels <strong>and</strong> can survive long dry spells by remaining in<br />

burrows. Red swamp crayfish are omnivorous, feeding on aquatic plants, snails, insects, fish<br />

<strong>and</strong> amphibian eggs <strong>and</strong> young. It seems that this alien species tend to reduce amphibians<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Silva et al. 2010. Using Business <strong>and</strong> Biodiversity to put Conservation into practice<br />

310<br />

populations in the Caridade stream (the main watercourse, which also floods the dam). This<br />

problem can partially be solved in time as the european otter (Lutra lutra) <strong>and</strong> some waterfowl<br />

are increasingly feeding on the crayfish. The ecological flow regime recognizes that flow<br />

magnitude, duration, frequency, timing, <strong>and</strong> predictability must be incorporated into any flow<br />

management strategy (Suen, 2005). So, changes in hydrologic trends, in concern to the<br />

continuing decline in the water levels in Caridade stream were solved through ecological flow<br />

management.<br />

Five RELAPE species were found in the areas surveyed (Ophrys tenthredinifera,<br />

Narcissus bulbocodium, Linaria spartea, Narcissus jonquilla <strong>and</strong> Genista polyanthos). Caridade<br />

vegetation is dominated by Fraxinus angustifolia Vahl galleries <strong>and</strong> woodl<strong>and</strong>s (Ficario<br />

ranunculoidis – Fraxinetum angustifoliae). Beneath <strong>and</strong> next the mature canopy it also can be<br />

described a high scrubl<strong>and</strong> community of Rubo ulmifolii – Nerietum ole<strong>and</strong>ri. Aquatic<br />

vegetation is also present, with Chara sp., Ranunculus peltatus Schrank <strong>and</strong> Callitriche<br />

stagnalis Scop. communities. These aquatic communities are extremely important to amphibian<br />

population, as refuge areas.<br />

The dam area is the most important for aquatic birds, despite the poor presence of vegetation<br />

communities. There are some communities of Typha angustifolia L. <strong>and</strong> Phragmites australis<br />

(Cav.) Trin. in the South East of the dam (Typho angustifoliae – Phragmitetum australis), but<br />

all the remnant area hasn’t vegetation. So the main action is to increase patches of vegetation by<br />

rooting some adequate plants.<br />

Synthetic table of the Ficario ranunculoidis – Fraxinetum angustifoliae Rivas‐Martínez &<br />

Costa in Rivas‐Martínez, Costa, Castroviejo & E. Valdés 1980, at Caridade stream, with low<br />

inclination, N, 100 m². Characteristics: Fraxinus angustifolia 2, Crataegus monogyna 2,<br />

Ranunculus ficaria 1, Aristolochia paucinervis 1, Scilla peruviana +; Companions: Pyrus<br />

bourgeana 2, Nerium ole<strong>and</strong>er 1, Asparagus acutifolius 1, Smyrnium olusatrum 1, Gyn<strong>and</strong>riris<br />

sysrinchium 1, Dactylis hispanica 1, Geranium molle 1, Ornithogalum baeticum +, Narcissus<br />

bulbocodium +, Ophrys tenthredinifera +.<br />

Rosemary scrubl<strong>and</strong><br />

Scattered throughout the Esporão l<strong>and</strong>scape, many patches with rosemary scrubl<strong>and</strong> which have<br />

importance to pollinator ecosystems service can be found. These also provide stepping-stone<br />

corridors to many species of invertebrates, reptiles, amphibians <strong>and</strong> birds.<br />

The flowers of rosemary produce pollen <strong>and</strong> nectar to attract insect populations so the<br />

maintenance of scrub dominated vegetative complex within Lav<strong>and</strong>ula stoechas L. is important.<br />

In order to effectively conserve these scrubl<strong>and</strong> areas, management of HCVA must restore <strong>and</strong><br />

maintain rosemary habitats. To achieve this, l<strong>and</strong> managers must minimize grazing intensity.<br />

Further work will include the measure of pollination services, so it is important to assess the<br />

rosemary scrubl<strong>and</strong> habitat to measure its ecosytemic value. Also, the maintained of these<br />

habitats in Esporão l<strong>and</strong>scape tend to safeguard <strong>and</strong> enhance pollination function in order to<br />

ensure effective pollination of wild plants (Potts, 2006), specially orchids species that are<br />

pollination limited, may also be enhanced by maintain rosermay patches in l<strong>and</strong>scape.<br />

4. Discussion<br />

The Esporão estate habitats <strong>and</strong> l<strong>and</strong>scapes are quite diverse, with important high conservation<br />

value species despite some forest areas degradation. We have assessed that dynamic plant<br />

communities are present, which are intensively interconnected in ecotone regions defining<br />

different ecological corridors. Also, Biodiversity Action Plans (BAPs) are, to some extent, a<br />

valid tool for l<strong>and</strong> owners / managers, by facilitating the identification <strong>and</strong> mapping of High<br />

Conservation Value Areas (HCVA) <strong>and</strong> important plant communities/assemblages that include<br />

rare, endemic, localized, threatened <strong>and</strong>/or endangered plant species, thus allowing the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Silva et al. 2010. Using Business <strong>and</strong> Biodiversity to put Conservation into practice<br />

311<br />

definition of specific management actions for each conservation area/value identified. Since<br />

2008 the following issues are ongoing: ecological evaluation of the vineyards natural enemies<br />

<strong>and</strong> ecological infrastructures; halting all predator control programs; reformulation of the<br />

forestry management plan; reduction of the area of pine plantations <strong>and</strong> implementation of<br />

biodiverse forest patches. In respect to agricultural l<strong>and</strong>scape, olive <strong>and</strong> vineyards they are<br />

being redesigned to achieve the following goals: optimize productivity, facilitate management<br />

<strong>and</strong> operations, prevent soil erosion <strong>and</strong> reduce the use of pesticides.<br />

The use of plants <strong>and</strong> plant communities as indicators for l<strong>and</strong> planning to describe <strong>and</strong> evaluate<br />

Esporão Estate has proven useful since it for HCVA protection. As well as the use of<br />

phytosociology can be used to describe the l<strong>and</strong>scape units, it also is important in what takes<br />

concern to ecological restoration. If habitats can be maintained then also species conservation is<br />

facilitated. The most represented phytosociological classes are: the Holm oak woodl<strong>and</strong>s Pyro<br />

bourgaeanae-Quercetum rotundifoliae <strong>and</strong> at streamside woodl<strong>and</strong> communities of Ficario<br />

ranunculoidis – Fraxinetum angustifoliae.<br />

Future research will focus on the validation of management effectiveness <strong>and</strong> on community<br />

dynamics.<br />

References<br />

Braun-Blanquet, J., 1979. Fitossociologia – Bases para el estudio de las comunidades<br />

vegetales. H. Blume ediciones. Madrid.<br />

Castroviejo, S. et. al., 1986-1997. Flora Iberica. Madrid. Real Jardín Bot. Madrid:1,2,3,4,5,6,8.<br />

Costa, J.C.; Aguiar, C.; Capelo, J.H.; Lousã, M. & Neto, C., 1998. Biogeografia de Portugal<br />

Continental. Quercetea, 0.<br />

Coutinho, A.X.; 1939. Flora de Portugal. Bertr<strong>and</strong>. Lisboa.<br />

Franco, J.A., 1971-1984. Nova Flora de Portugal (Continente e Açores). <strong>Escola</strong>r Editora,<br />

Lisboa.Vol. I e II.<br />

Franco, J.A. <strong>and</strong> Rocha Afonso, M.L., 1994-1998-2003. Nova Flora de Portugal (Continente e<br />

Açores). <strong>Escola</strong>r Editora, Lisboa. Vol. III, fasc. 1,2,3.<br />

Géhu, J.M. <strong>and</strong> Rivas-Martínez, S., 1981. Notions fondamentales de Phytosociologie in<br />

Syntaxonomie. J. Cramer. Vaduz.<br />

Lengyel, S., Kobler, A., Kutnar, L., Framstad, E., Henry, P., Babij, V., Gruber, B., Schmeller,<br />

D. & Henle, K., 2008. A review <strong>and</strong> a framework for the integration of biodiversity<br />

monitoring at the habitat level. Biodiversity Conservation, 17: 3341-3356.<br />

Loidi, J., 1994. Phytosociology applied to nature conservation <strong>and</strong> l<strong>and</strong> management. 35th<br />

Symposium IAVS. East China Normal Univ. Press, 17-30.<br />

Pereira, M. & Costa, J.C., 2004. Sintaxonomia das classes fitossociológicas em Portugal<br />

Continental. Universidade de Évora, Évora.<br />

Potts, S.; Petanidoub, T.; Roberts, S.; Toolec, C.; Hulbertd, A. & Willmerd, P., 2006. Plantpollinator<br />

biodiversity <strong>and</strong> pollination services in a complex Mediterranean l<strong>and</strong>scape.<br />

Biological Conservation, 129: 519-529<br />

Rivas-Martínez, S.; Fern<strong>and</strong>éz-González, F.; Loidi, J.; Lousã, M. & Penas, A., 2001.<br />

Syntaxonomical checklist of Vascular Plant Communities of Spain <strong>and</strong> Portugal to<br />

association level. Itinera Geobotanica 14.<br />

Rivas-Martínez, S.; Diaz, T.E.; Fern<strong>and</strong>éz-González, F.; Izco, J.; Loidi, J.; Lousã, M. & Penas,<br />

A., 2002 a, b. Syntaxonomical checklist (2001) of Vascular Plant Communities of Spain<br />

<strong>and</strong> Portugal to association level. Itinera Geobotanica 15 (1) (2).<br />

Suen, J.; Eheart, J. & Herricks, E., 2005. Integrating Ecological Flow Regimes in Water<br />

Resources Management Using Multiobjective Analysis. Proceedings of World Water <strong>and</strong><br />

Environmental Resources Congress.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.N. Terra & R.F dos Santos 2010. Legal efficiency <strong>and</strong> cumulative effects in environmentally protected area<br />

312<br />

Legal efficiency <strong>and</strong> cumulative effects in environmentally protected<br />

area<br />

Talita Nogueira Terra * & Rozely Ferreira dos Santos<br />

LAPLA, Dept. of Water Resources, Energy <strong>and</strong> Environmental, UNICAMP, Av. Albert<br />

Einstein, 951, 13083-970, Campinas, SP, Brazil<br />

Abstract<br />

Decisions must be taken according to the impacts observed over a temporal sequence of<br />

human action. But there is another side to consider - the efficiency of the legal action after<br />

decree. In summary, l<strong>and</strong>scapes are complexes <strong>and</strong> involve a system of multiple overlapping of<br />

l<strong>and</strong> use <strong>and</strong> legal decisions. The objective of this study is interpret the chains of cumulative<br />

effect four decades to estimate the cumulative of these impacts in the future <strong>and</strong> relate the<br />

results of these chains in the face of environmental laws in force in that period. The study area<br />

was composed by three regions adjacent one each other: the Sustainable Development Reserve<br />

of Despraiado, the Ecological Station of Juréia-Itatins <strong>and</strong> the buffer zone. The cumulative<br />

effects were inferred by constructing scenarios of past <strong>and</strong> present. This strategy should explain<br />

state changes along the series of l<strong>and</strong> use <strong>and</strong> simulate the changes in the coming years.<br />

Keywords: cumulative effects, protected area, legal action<br />

1. Introduction<br />

The increase in human population causes changes in l<strong>and</strong> use, affecting the properties of the<br />

l<strong>and</strong>scape. Certain human actions that commonly occur in sequence, as deforestation, burning<br />

<strong>and</strong> opening of roads, can cause significant changes on the natural processes (Thomaziello,<br />

2007). The acceleration of the processes of environmental degradation <strong>and</strong> the cumulative<br />

effects occur in the concentration of humans in a physical space <strong>and</strong> the expansion of l<strong>and</strong> use<br />

which is related to the increase of human population, with increasing dem<strong>and</strong> for consumer<br />

goods <strong>and</strong> increasing consumption of materials <strong>and</strong> energy (Coelho, 2001 <strong>and</strong> Chen, 2006).<br />

The cumulative effects are the changes in the environment caused by the combination of<br />

human actions in the past, present <strong>and</strong> future (Hegmann et al. 1999). Human activities have<br />

their effects accumulated when the second disturbance occurs in the same location in which the<br />

first one occurred, while the ecosystem is recovering from the effects of the first one (NEPA,<br />

1997). The cumulative effects are currently resulting of multiple activities in the territory that<br />

persist for a long time. Therefore, the attribute space becomes important when integrated with<br />

variable time. In summary, l<strong>and</strong>scapes are complex <strong>and</strong> involve a system of multiple<br />

overlapping in space <strong>and</strong> time <strong>and</strong> are subject to multi-use over a number of anthropogenic<br />

changes (MacDonald, 2000). It is common that cumulative effects are greater than actually<br />

seem to be, because the summation of individual impacts can be interacting or multiply the<br />

effects (Halpern et al., 2008), in the other words they may have actions of synergism or<br />

addition. Therefore they are difficult to mensure.<br />

Santos & Santos (2008 a, b), who studied the region's agricultural Andradina (Sao<br />

Paulo), the main observed change in l<strong>and</strong> use was strongly associated with the decline of<br />

agriculture, the loss of plantation areas due the expansion of pastures <strong>and</strong> the increase soil areas<br />

with difficult recovery. They applied measures for change of l<strong>and</strong> use or destination of the l<strong>and</strong><br />

along thirty years, which coincided with significant differences about historical changes, linked<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.N. Terra & R.F dos Santos 2010. Legal efficiency <strong>and</strong> cumulative effects in environmentally protected area<br />

313<br />

to political <strong>and</strong> government actions. In these studies, the spatial measures were made by<br />

applying the rates of change <strong>and</strong> to better underst<strong>and</strong> them, the historical information of the<br />

region were associated with the own perception of the community about the facts.<br />

The implementation of Conservation Areas (CA) in Brazil occurs on regions already<br />

occupied with a history of impacts, traces left by human action. When a CA is established, it is<br />

expected that its impacts are at least reduced through management actions, hoping that the<br />

reversibility of the process occurs toward conservation. Therefore, decisions must be taken<br />

according to the impacts observed over a temporal sequence of human action. Should be<br />

considered, for example, that an Ecological Station cannot have human occupation, according to<br />

SNUC (National System of Conservation Units), however it is inevitable that the area bring the<br />

impacts that have been made by human occupation in the past <strong>and</strong> it reflect in the present<br />

observed. In addition, each CA has its own characteristics, with specific objectives to achieve a<br />

degree of conservation. So, the actions of management will be different between an Ecological<br />

Station <strong>and</strong> a Sustainable Development Reserve. But there is another side to consider - the<br />

efficiency of the legal action after decree. In summary, l<strong>and</strong>scapes are complexes <strong>and</strong> involve a<br />

system of multiple overlapping of l<strong>and</strong> use <strong>and</strong> legal decisions. Therefore, the objective of this<br />

study was interpret the effect chains for two decades to estimate accumulation of these impacts<br />

in the past <strong>and</strong> present <strong>and</strong> relate the results against environmental laws in force in that period.<br />

2. Method<br />

The study area is located at the southern of Sao Paulo state, <strong>and</strong> covers parts of three<br />

distinct adjacent regions: the oldest Sustainable Development Reserve of Despraiado (SDRD),<br />

the Juréia-Itatins State Ecological Station (JISES) <strong>and</strong> buffer zone (BZ) adjacent to these two<br />

conservation areas (Figure 1). These three regions have an area of about 100ha each one.<br />

The impacts were inferred by the analysis of overlapping l<strong>and</strong> use from the scenarios of<br />

past. The overlapping was made by the tool spatial statistics available in the software Arc Gis<br />

9.2.<br />

The scenarios were constructed from the photointerpretation of 11 types of l<strong>and</strong> use,<br />

namely: agriculture; grazing fields; fields like lawns/gardens <strong>and</strong> grounds around the building<br />

originating fields; fields of infrastructure; construction; water tanks; Tropical Rainforest<br />

Secondary Initial; Tropical Rainforest Secondary Medium; Tropical Rainforest in areas with<br />

bananas; crop rotation/ab<strong>and</strong>oned areas; access roads. The scenarios were constructed for 1980<br />

<strong>and</strong> the 2007. The aerial photograph of 1980 was acquired in digital format arising directly from<br />

the roll of photographic film, scanned in photogrammetric scanner Vexcel Ultrascan 5000, with<br />

1200 dpi resolution. For the scenario of 2007 was used a satellite image of World View with<br />

spatial resolution of 0.5 PAN (provided by the Foundation <strong>Forest</strong>-Brazil).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.N. Terra & R.F dos Santos 2010. Legal efficiency <strong>and</strong> cumulative effects in environmentally protected area<br />

314<br />

Figure 1: Localization of study area.<br />

3. Result <strong>and</strong> Discussion<br />

In 1980, the predominant element in the both conservation areas was the Tropical<br />

Rainforest both in Sustainable Development Reserve of Despraiado like in Juréia-Itatins State<br />

Ecological Station (Figure 2). After 27 years of the federal act of the JISES was observed that<br />

the l<strong>and</strong> uses increased <strong>and</strong> vegetation cover decreased dramatically. In this unit left 34% of the<br />

Tropical Rainforest <strong>and</strong> had an increased by 47% in agriculture <strong>and</strong> 83% the banana inserted<br />

into the forest. The data indicate that most of the forested areas have been transformed by man<br />

in uses, such as for crops, pastures, buildings <strong>and</strong> access roads.<br />

On the other h<strong>and</strong>, the agriculture gained area in the JISES <strong>and</strong> in the SDRD. Observed that<br />

in SDRD, the area of Tropical Rainforest Secondary Medium lost area for the agriculture, for<br />

the Tropical Rainforest Secondary Initial <strong>and</strong> for the forest with bananas. The fact is this area<br />

has been suffered strong influence of human activities <strong>and</strong> the legal acts have not been respected<br />

(Figure 3).<br />

The buffer zone is an area that has strong impact since 1980. The probable reasons are the<br />

little distance from an important highway, the road Padre Manoel da Nóbrega. Moreover, it is<br />

surrounded by many little cities. In 1980 can be observed a big area for agriculture like as<br />

banana plantation <strong>and</strong> in 2007 the area decreased. In terms of gains <strong>and</strong> losses relating, this area<br />

presents a lose in terms of Tropical Rainforest but gained in terms of uses that caused impacts in<br />

the area.<br />

In this way, we can conclude that the main goal of the creation of conservation areas has not<br />

been reached, in other words, recovery, conservation <strong>and</strong> protection of the forest, which was<br />

established by National System of Conservation Units.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.N. Terra & R.F dos Santos 2010. Legal efficiency <strong>and</strong> cumulative effects in environmentally protected area<br />

315<br />

Figure 2: Maps of l<strong>and</strong>-use <strong>and</strong> l<strong>and</strong>-cover change in 1980 <strong>and</strong> 2007, scale 1:35.000.<br />

Figure 3: Relative change in the use <strong>and</strong> occupancy of the study area between 1980 <strong>and</strong> 2007.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T.N. Terra & R.F dos Santos 2010. Legal efficiency <strong>and</strong> cumulative effects in environmentally protected area<br />

316<br />

References<br />

Chen, C. - C., 2006. Development of a framework for sustainable uses of resources: More paper<br />

<strong>and</strong> less plastics Environment International, 32: 478-486.<br />

Coelho, M.C.N., 2001. Impactos Ambientais em Áreas Urbanas. In: Antônio José Teixeira<br />

Guerra e S<strong>and</strong>ra Baptista da Cunha. (Org.). Impactos Ambientais Urbanos no Brasil. 1ª<br />

ed. Rio de Janeiro, Brazil: Bertr<strong>and</strong> Brasil: 19-45.<br />

Halpern, S.B. et al. 2008. Managing for cumulative impacts in ecosystem-based management<br />

through ocean zoning. Ocean & Coastal Management, 51: 203-211.<br />

Hegmann, G., Cocklin, C., Creasey, R., Dupuis, S., Kennedy, A., Kingsley, L., Ross, W.,<br />

Spaling, H. <strong>and</strong> Stalker, D., 1999. Cumulative effects assessment practitioners guide.<br />

Hull, Quebec: AXYS Environmental Consulting Limited, CEA Working Group for the<br />

Canadian Environmental Assessment Agency.<br />

MacDonald, L.H., 2000. Evaluating <strong>and</strong> Managing Cumulative Effects: Process <strong>and</strong><br />

Constraints. Environmental Management, 26: 3: 299-315.<br />

NEPA. 1997. Considering cumulative effects under the National Environmental Policy Act.<br />

National Ennvironmental Politic Act.<br />

Santos, M.A.; Santos, R.F. Construção de cenários por análises temporais e métricas espaciais.<br />

Revista do Instituto Florestal, 2008.<br />

Santos, M.A.; Santos, R.F., 2008. Aplicación de índices de cambios para evaluación de las<br />

auteraciones en el uso de las tierras. Investigaciones Geográficas. Instituto de Geografía.<br />

Universidad Nacional Autónoma de México. “No Prelo”.<br />

Thomaziello, S., 2007. Usos da terra e sua influência sobre a qualidade ambiental. In:<br />

Vulnerabilidade Ambiental: Desastres naturais ou fenômenos induzidos Brasilia: MMA:<br />

23-38.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Tetetla-Rangel et al. 2010. Relationships among l<strong>and</strong>scape structure, climate <strong>and</strong> rare woody species richness<br />

317<br />

Relationships among l<strong>and</strong>scape structure, climate <strong>and</strong> rare woody<br />

species richness in the tropical forests of the Yucatan Peninsula,<br />

Mexico<br />

Erika Tetetla-Rangel * , J. Luis Hernández-Stefanoni & Juan Manuel Dupuy<br />

Centro de Investigación Científica de Yucatán A.C. Unidad de Recursos Naturales,<br />

Calle 43 # 130. Chuburná de Hidalgo. C.P. 97200, Mérida, Yucatán, México<br />

Abstract<br />

Two key concerns about the tropical forests of the Yucatan Peninsula are: the effects of l<strong>and</strong>-use<br />

<strong>and</strong> climate change on biodiversity, <strong>and</strong> the scant knowledge about rare plants species, which<br />

are most vulnerable to these changes. We assessed the association of rare woody species<br />

richness of tropical forests of the Yucatan Peninsula, classified in three levels of rarity (low,<br />

medium <strong>and</strong> high), with l<strong>and</strong>scape structure, climate <strong>and</strong> spatial dependence. We estimated the<br />

number of rare species within 25km 2 l<strong>and</strong>scape units, which were characterized using climate,<br />

altitude <strong>and</strong> l<strong>and</strong>scape configuration. Principal Neighbor Coordinate of Matrices (PCNM) <strong>and</strong><br />

regression analyses were performed to decompose variation of rare species richness into<br />

environmental <strong>and</strong> spatial components. Space was the most important variable related to the<br />

number of rare species in each of the three levels of rarity. These results suggest that dispersal<br />

limitation or other environmental variables not considered drive richness of rare woody species.<br />

Keywords: L<strong>and</strong>scape structure; rare woody plant species; spatial dependence; Tropical forest.<br />

1. Introduction<br />

Tropical forests of the Yucatan Peninsula have been converted to other l<strong>and</strong> uses with unknown<br />

effects on biodiversity. The floristic diversity of these ecosystems is mainly composed of rare<br />

species –those having small populations, high habitat specificity <strong>and</strong>/or restricted distribution–<br />

which are most vulnerable to extinction due to transformation <strong>and</strong> loss of habitat (Rabinowitz<br />

1981).<br />

To preserve rare plant species we need to know where they occur <strong>and</strong> what factors determine<br />

their presence <strong>and</strong> abundance. Many factors affect the distribution of rare species, such as<br />

historic events, climate, biotic interactions <strong>and</strong> spatial structure. Several studies have<br />

demonstrated that environmental conditions (climatic, topographic <strong>and</strong> soil factors) constrain<br />

the abundance or restrict the distribution range of these species. However, the factors that<br />

determine species distribution vary depending on the way in which rarity is defined (Gaston<br />

1994).<br />

Several studies report a significant relationship between l<strong>and</strong>scape configuration <strong>and</strong> plant<br />

species diversity (Bascompte <strong>and</strong> Rodríguez 2001; Hern<strong>and</strong>ez-Stefanoni, 2005). However, few<br />

studies have related l<strong>and</strong>scape structure <strong>and</strong> habitat heterogeneity to rare plant species, due to<br />

difficulties involved in obtaining field information of this group of species (Luoto 2000).<br />

* Corresponding author.<br />

Email address: tetetla@cicy.mx<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Tetetla-Rangel et al. 2010. Relationships among l<strong>and</strong>scape structure, climate <strong>and</strong> rare woody species richness<br />

318<br />

The aim of this study was to evaluate the association of rare woody species richness at different<br />

levels of rarity with l<strong>and</strong>scape structure, climate <strong>and</strong> spatial dependence, to generate scientific<br />

knowledge necessary to promote the conservation <strong>and</strong> management of biodiversity of tropical<br />

forests.<br />

2. Methodology<br />

2.1 Study area<br />

The study area is located at the Yucatan peninsula in SE Mexico (17° 50´ to 21° 30´ north<br />

latitude <strong>and</strong> 87° 00´ to 91° 00´ east longitude), covering a total area of 141,523 km 2 (Figure 1).<br />

The peninsula is constituted almost entirely of limestone Karsts from the Eocene. Around 95%<br />

of the territory consists of flat lowl<strong>and</strong>s < 100 m above sea level, but some hills reach 275<br />

meters (Ferrusquía-Villafranca 1993). Seven types of tropical forests, in different stages of<br />

secession, cover the peninsula (Flores <strong>and</strong> Carvajal 1994; Durán et al. 1999) <strong>and</strong> their<br />

distribution patterns are determined by precipitation <strong>and</strong> soil type (Carnevali et al. 2003).<br />

.<br />

Mexico<br />

200 0 200 400 km<br />

Rarewoody species records<br />

<strong>L<strong>and</strong>scapes</strong>25km²<br />

Main citys<br />

#<br />

#<br />

# #<br />

#·<br />

Campeche<br />

##<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

#<br />

## #<br />

#<br />

#<br />

# #<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

#<br />

#<br />

Merida<br />

#<br />

# ##<br />

#<br />

# #·<br />

#<br />

#<br />

#<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

##<br />

##<br />

##<br />

#<br />

#<br />

# # #<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

##<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

##<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

## #<br />

#<br />

## #<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

##<br />

#<br />

#<br />

# #<br />

##<br />

#<br />

#<br />

#<br />

#<br />

# #<br />

Cancun<br />

#<br />

#·<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

##<br />

#<br />

#<br />

#<br />

#<br />

#<br />

##<br />

# # #<br />

#<br />

#<br />

##<br />

##<br />

#<br />

##<br />

N<br />

30 0 30 60 km<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

##<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

##<br />

#<br />

#<br />

#<br />

#<br />

##<br />

#<br />

#<br />

#<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

#<br />

#<br />

##<br />

#<br />

##<br />

#<br />

#<br />

#<br />

##<br />

#<br />

# # #<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

# Chetumal<br />

#<br />

#·<br />

##<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

# #<br />

#<br />

#<br />

#<br />

##<br />

#<br />

#<br />

#<br />

#<br />

#<br />

#<br />

Figure 1: Study area <strong>and</strong> rare woody species records within 25 km 2 l<strong>and</strong>scapes.<br />

2.1.2 Rare woody species richness<br />

Using herbarium records (Centro de Investigación Científica de Yucatán CICY) we identified<br />

rare woody species, which were classified in three levels of rarity – low, medium <strong>and</strong> high–<br />

according of their frequency, habitat specificity <strong>and</strong> range of potential distribution (modelled in<br />

DOMAIN, Carpenter et al. 1993). To calculate rare woody species richness we generated a 25<br />

km 2 grid (l<strong>and</strong>scapes) covering the whole study area (Figure 1).<br />

2.1.3 Environmental variables<br />

Three groups of environmental variables were considered as explanatory variables in the<br />

regression models in those l<strong>and</strong>scapes containing at least one rare species registered: (1) climate,<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Tetetla-Rangel et al. 2010. Relationships among l<strong>and</strong>scape structure, climate <strong>and</strong> rare woody species richness<br />

319<br />

(2) topography <strong>and</strong> (3) l<strong>and</strong>scape configuration. Several digital climatic maps were interpolated<br />

using kriging <strong>and</strong> the software GS+ (Version 9) to extract information of the average, st<strong>and</strong>ard<br />

deviation <strong>and</strong> coefficient of variation of annual precipitation as well as mean, maximum <strong>and</strong><br />

minimum annual temperature. We also extracted the altitude from a digital elevation model<br />

(DEM). We calculated the following l<strong>and</strong>scape metrics: number of patches (NP), patch density<br />

(PD), largest patch index (LPI), edge density (ED), l<strong>and</strong>scape shape index (LSI), total edge<br />

contrast index (TECI), patch richness (PR), Shannon’s diversity index (SHDI) <strong>and</strong> Simpson’s<br />

diversity index (SIDI) using FRAGSTATS 3.0 (McGarical et al. 2002).<br />

2.1.4 Data analyses<br />

To determine the spatial structure of l<strong>and</strong>scape locations <strong>and</strong> include this as a variable to predict<br />

rare woody species richness, we used principal coordinate of neighbor matrices (PCNM, Bocard<br />

et al. 2004). To partition the variability into environmental <strong>and</strong> space components, we used<br />

three sets of multiple linear regression models using different independent variables: (1) PCNM<br />

vectors, (2) environmental variables, <strong>and</strong> (3) all selected variables in (1) <strong>and</strong> ( 2).<br />

3. Results<br />

To divide rare species into groups, we first selected those species representing the 25th<br />

percentile of frequency distribution of herbarium records to identify species having low<br />

frequency. We then assigned these species into 2 categories of habitat specificity according to<br />

information from the CICY <strong>and</strong> Missouri herbariums: high habitat specificity (species restricted<br />

to 1 or 2 vegetation types) <strong>and</strong> low specificity (species reported in > 2 vegetation types). Finally,<br />

we used the 25th percentile of the modelled area of potential distribution to assign species into<br />

narrow <strong>and</strong> broad classes (Table 1).<br />

Table 1: Criteria used to define rarity levels of woody species.<br />

Frequency Specificity Distribution Rarity level<br />

Low (< 5 records) Low (3 - 6 habitat types) Wide (> 9886.25 km²) Low<br />

Low (< 5 records) High (1 - 2 habitat types) Wide (> 9886.25 km²) Medium<br />

Low (< 5 records) Low (3 - 6 habitat types) Narrow (< 9886.25 km²) Medium<br />

Low (< 5 records) High (1 - 2 habitat types) Narrow (< 9886.25 km²) High<br />

We found 242 rare woody plant species (133 trees, 85 shrubs <strong>and</strong> 31 lianas) out of the 955<br />

woody species registered in the herbarium for the Yucatan peninsula. Due to the lack of data to<br />

model the potential distribution of rare woody species, we only used 138 species (76 trees, 45<br />

shrubs <strong>and</strong> 17 lianas) to generate the potential distribution maps of these rare species, since the<br />

remaining 104 species had only 1 herbarium sample. We found 18 species belonging to the low<br />

level of rarity, <strong>and</strong> 34 <strong>and</strong> 86 belonging to the medium <strong>and</strong> high levels, respectively.<br />

Total variation in rare species richness explained by environmental variables ranged from 3 to<br />

16% (Table 2). The medium level of rarity had the highest percentage of variation explained <strong>and</strong><br />

was associated with the highest number of environmental variables. Species richness was<br />

negatively associated with maximum temperature <strong>and</strong> positively related to precipitation.<br />

Relationships involving TECI <strong>and</strong> LSI were negative, indicating that rare species richness<br />

decreases as the contrast of patches increases <strong>and</strong> as the shape of a l<strong>and</strong>scape becomes more<br />

irregular. On the other h<strong>and</strong>, the number of patches <strong>and</strong> edge density were positively related to<br />

rare species of this level.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Tetetla-Rangel et al. 2010. Relationships among l<strong>and</strong>scape structure, climate <strong>and</strong> rare woody species richness<br />

320<br />

On the other h<strong>and</strong>, variation partitioning revealed that spatial dependence was the variable most<br />

strongly associated with rare species richness; the amount of variation explained by this variable<br />

ranged from 13 to 17% for the different levels of rarity (Figure 2). Environmental variables<br />

explained a higher percentage of variation for the low <strong>and</strong> medium levels of rarity, compared to<br />

the high level <strong>and</strong> to all rare species.<br />

Table 2: Regression st<strong>and</strong>ardized coefficients <strong>and</strong> partial multiple correlations for predicting rare woody<br />

species richness from environmental attributes.<br />

CLIMATE AND<br />

Level of rarity<br />

LANDSCAPE STRUCTURE All rare spp. Low Medium High<br />

(R 2 = 0.05)* (R 2 = 0.08)* (R 2 = 0.16)* (R 2 = 0.03)*<br />

MEAN PRECIPITATION -0.22 (0.05) 0.17 (0.03) 0.17 (0.03)<br />

MEAN TEMPERATURE MAXIMUM -0.17 (0.03) -0.20 (0.03)<br />

SD ALTITUDE -0.17 (0.03)<br />

TECI 0.25 (0.03) -0.27 (0.02)<br />

LSI -0.20 (0.02) -0.46 (0.03)<br />

ED 0.58 (0.03)<br />

NP 0.24 (0.03)<br />

* Coefficient of determination for the model.<br />

n<br />

tio<br />

v<br />

a<br />

ria<br />

e<br />

d<br />

in<br />

e<br />

x<br />

p<br />

la<br />

o<br />

f<br />

g<br />

e<br />

e<br />

n<br />

ta<br />

e<br />

rc<br />

P<br />

25<br />

20<br />

15<br />

10<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

5<br />

0<br />

a)<br />

c)<br />

Environmental Space Shared Total<br />

Environmental Space Shared Total<br />

b)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Environmental Space Shared Total<br />

d)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Environmental Space Shared Total<br />

Components<br />

Figure 2: Partitioning of the variation of rare species richness by environmental <strong>and</strong> space components.<br />

All species (a), low level of rarity (b), medium level of rarity (c), <strong>and</strong> high level of rarity (d).<br />

4. Discussion<br />

The most important variable explaining variation in rare woody species richness was spatial<br />

dependence (Figure 2). This result suggests that the distribution of these species is strongly<br />

affected by dispersal limitation (Hubbell 2001). Alternatively, rare species distribution may be<br />

influenced by other environmental variables operating at local scales, such as physical <strong>and</strong><br />

chemical soil properties, <strong>and</strong> microclimate (Jones et al. 2008; Lindo <strong>and</strong> Winchester 2009).<br />

However, the relative contribution of spatial <strong>and</strong> environmental components varied among the<br />

different levels of rarity. We found that as the level of rarity increases, the amount of variation<br />

explained by environmental variables decreases. This suggests that dispersal limitation or local<br />

environmental factors may be most important for species with the highest level of rarity, that is,<br />

those with low frequency, high levels of habitat specificity, <strong>and</strong> restricted distribution. Since the<br />

majority of rare species studied belonged to the highest level of rarity, the patterns shown by all<br />

rare species were very similar to those of species with the highest level of rarity.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Tetetla-Rangel et al. 2010. Relationships among l<strong>and</strong>scape structure, climate <strong>and</strong> rare woody species richness<br />

321<br />

On the other h<strong>and</strong>, we suspect that l<strong>and</strong>scape structure <strong>and</strong> climate are more relevant for species<br />

with low <strong>and</strong> medium levels of rarity. In particular, species with medium level of rarity were<br />

positively associated with precipitation <strong>and</strong> negatively associated with maximum temperature.<br />

This result suggests that rare species tend to be more common in mesic environments. For<br />

example, Gaston (1994) reported that precipitation is an excellent predictor variable of patterns<br />

of distribution of rare species. In this study the highest number of rare species was found in the<br />

south of the Peninsula, according to the precipitation gradient from the North-West (lowest) to<br />

the South-East (highest).<br />

We found apparently contradictory results regarding the relationship between l<strong>and</strong>scape<br />

structure <strong>and</strong> the number of species with medium level of rarity (Table 2). On the one h<strong>and</strong>, rare<br />

species richness decreased as the contrast among patch types increased <strong>and</strong> as their shape was<br />

more irregular, suggesting that l<strong>and</strong>scape heterogeneity has a negative effect on rare species<br />

richness. On the other h<strong>and</strong>, edge density <strong>and</strong> number of patches were positively related to rare<br />

species richness, suggesting an increase in the number of species in more heterogeneous<br />

l<strong>and</strong>scapes. These results may indicate that small disturbances promote spatial variation of<br />

patches, which create habitat diversity allowing an increase in the number of species. However,<br />

large amounts of disturbance increase the contrast of patches, indicating a major degree of<br />

fragmentation that negatively impacts rare plant species (Honnay et al. 2003).<br />

Finally, our results indicate that most of the variation in rare woody species richness was<br />

unexplained indicating the need for more detailed studies to elucidate the factors determining<br />

rare species distribution <strong>and</strong> richness. However, we did find some relationships between<br />

environmental factors studied <strong>and</strong> rare woody species richness, which can contribute to the<br />

conservation <strong>and</strong> management of these species.<br />

References<br />

Bascompte, J. <strong>and</strong> Rodríguez, M.A., 2001. Habitat patchiness <strong>and</strong> plant species richness.<br />

Ecology Letters, 4: 417-420.<br />

Bocard, D., Legendre, P., Avois-Jacquet, C. <strong>and</strong> Tuomisto, H ., 2004. Dissecting the spatial<br />

structure of ecological data at multiple scales. Ecology, 85:1826-1832.<br />

Carnevali, G., Ramírez, M.I. <strong>and</strong> González-Iturbe, J.A., 2003. Flora y Vegetación de la<br />

Península de Yucatán. In: Colunga-GarcíaMarín, P. <strong>and</strong> Savedra, L.A (Eds.). Naturaleza<br />

y sociedad en el área maya, pasado presente y futuro. Academia Mexicana de las<br />

Ciencias/ Centro de Investigación Científica de Yucatán.<br />

Carpenter, G., Gillison, A.N. <strong>and</strong> Winter, J., 1993. DOMAIN: a flexible modeling procedure for<br />

mapping potencial distributions of plants <strong>and</strong> animals. Biodiversity <strong>and</strong> Conservation,<br />

2:667-680.<br />

Durán, G.R., González-Iturbe, A.J., Granados, C.J., Olmsted, C.I. <strong>and</strong> Tun, D.F., 1999.<br />

Vegetación. In: Garcia, A. <strong>and</strong> Cordoba, J. (Eds.). Atlas de procesos territoriales de<br />

Yucatán. 184-194.<br />

Ferrusquía-Villafranca, I., 1993. Geology of Mexico: A Synopsis. In: Ramamoorthy, P.T., Bye,<br />

R., Lot, A. <strong>and</strong> Fa, J. (Eds.). Biological Diversity of Mexico: Origins <strong>and</strong> Distribution.<br />

Oxford. New York.<br />

Flores, S.J <strong>and</strong> Carvajal, E.I., 1994. Descripción de los tipos de vegetación de la península de<br />

Yucatán. In: Gómes-Pompa, A., Flores, S.J., Sosa, O.V., Caballero, N.J., Chiang, C.F.,<br />

Diego, P.N., Ortíz, D.J.J., Martínez, A.M.A. <strong>and</strong> Guevara, S.S. (Eds.). Etnoflora<br />

yucatanense. Universidad Autónoma de Yucatán, México.<br />

Gaston, J. K., 1994. Causes of rarity. In: Usher, B.M. <strong>and</strong> DeAgelies, L.D (Eds.). Rarity.<br />

Chapman <strong>and</strong> Hall, London, UK<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Tetetla-Rangel et al. 2010. Relationships among l<strong>and</strong>scape structure, climate <strong>and</strong> rare woody species richness<br />

322<br />

Hernández-Stefanoni, J.L., 2005. Relationships between l<strong>and</strong>scape patterns <strong>and</strong> species richness<br />

of trees, shrubs <strong>and</strong> vines in a tropical forest. Plant Ecology, 179: 53-65.<br />

Honnay, O., Piessens, K., Van L<strong>and</strong>uyt, W., Hermy, M. <strong>and</strong> Guilinck, H. 2003. Satellite based<br />

l<strong>and</strong> use <strong>and</strong> l<strong>and</strong>scape complexity indices as predictors for regional plant species<br />

diversity. L<strong>and</strong>scape <strong>and</strong> Urban Planning. 63(4): 241-250.<br />

Hubbell, S. P. 2001. The unified neutral theory of biodiversity <strong>and</strong> biogeography. Princeton<br />

University Press, Princeton. 390 pp.<br />

Jones, M.M., Tuomisto, H. <strong>and</strong> Bocard, D., 2008. Explaining vatiation in tropical plant<br />

community composition: influence of environmental <strong>and</strong> spatial data quality. Community<br />

Ecology, 155: 593-604.<br />

Lindo, Z. <strong>and</strong> Winchester, N.N., 2009. Spatial <strong>and</strong> environmental factors contributing to<br />

patterns in arboreal <strong>and</strong> terrestrial oribatid mite diversity across spatial scales. Community<br />

Ecology, 160: 817-825.<br />

Luoto, M., 2000. Modelling of rare plant species richness by l<strong>and</strong>scape variables in an<br />

agricultura area in Finl<strong>and</strong>. Plant Ecology, 149: 157-168.<br />

McGarigal, K., Cushman, S.A., Neel, M.C. <strong>and</strong> Ene, E., 2002. FRAGSTAST: Spatial pattern<br />

analysis program for categorical maps. University of Massachusetts.<br />

Rabinowitz, D., 1981. Seven forms of rarity. In: Synge, H. (Ed.). The biological aspects of rare<br />

plant conservation. pp. 205-217. Engl<strong>and</strong>, Kew: The herbarium, Royal Botanic Gardens<br />

press: 189-217.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Tsibulnikova 2010. Economic estimations in using of the l<strong>and</strong>scapes planning<br />

323<br />

Economic estimations in planning of using <strong>and</strong> preservation of<br />

natural l<strong>and</strong>scapes<br />

M. Tsibulnikova<br />

Tomsk State University<br />

Abstract<br />

The very important task in preserving the natural l<strong>and</strong>scapes is to include the non-economical<br />

values in territory development management. Economical value definition of the natural<br />

resources based on market <strong>and</strong> non-market methods allow us to take into account ecological <strong>and</strong><br />

social aspects in economical estimation of the territory natural resources <strong>and</strong> to choose natural<br />

l<strong>and</strong>scapes using directions.<br />

Article is short presentation of results of the economic-geographical analysis of wildlife<br />

management in the Tomsk region.<br />

Taking into account a social factor, natural resources monetary estimations reflect economicalgeographical<br />

features of the territory, <strong>and</strong> indicate the stability of wildlife management <strong>and</strong> can<br />

be used in the natural l<strong>and</strong>scapes management.<br />

Keywords: economical estimation, natural capital, l<strong>and</strong>scape<br />

1. Introduction<br />

In accordance with the concept of sustainable development, striving for favourable living<br />

conditions <strong>and</strong> natural environment should be the basis of economic policy.<br />

Natural resources (mineral <strong>and</strong> raw material resources, forest <strong>and</strong> water, biological <strong>and</strong> other<br />

resources) are of great importance for social <strong>and</strong> economic development of Tomsk region.<br />

Tomsk region is located on West Siberian plain in the middle stream of River Ob’, its area<br />

making up about 316.9 thous<strong>and</strong> square kilometers. The climate of Tomsk region is continental<br />

due to its geographical location (temperate zone, in latitude 55-61° North). The average annual<br />

temperature is negative: –0.5°С to –3.5°С. At the same time the temperature may go up to more<br />

than +30°С in summer months, <strong>and</strong> go down to – 30°С in winter months. The population of the<br />

region is about 1.04 million people, including 67 % of city-dwellers.<br />

Tomsk region is an industrial region with highly developed technology, gas-<strong>and</strong>-oil production,<br />

petrochemistry, science, culture, <strong>and</strong> agriculture in southern areas.<br />

Total geological resources of oil <strong>and</strong> gas in Tomsk region are estimated at 5.4 billion ton<br />

st<strong>and</strong>ard units. (1 ton of oil is equal to 1 thous<strong>and</strong> cubic meter of gas <strong>and</strong> comprises 1 ton<br />

st<strong>and</strong>ard units of raw hydrocarbon).<br />

<strong>Forest</strong> covers about 61 % of the territory. Pine nut resources in average productivity years<br />

comprise up to 58.7 thous<strong>and</strong> ton. 16 sorts of mushrooms out of 72 sorts growing in the region<br />

are in use. Total regional biological resources of fruit <strong>and</strong> berry of all kinds comprise 58627 ton.<br />

59 sorts of medicinal herbs are found throughout Tomsk region. Raw material of 38 sorts of<br />

medicinal herbs is being prepared for the needs of drugstores, individuals, <strong>and</strong> market sales.<br />

An important contribution of the natural resources component to gross regional product, as well<br />

as potential increase of GRP owing to nature management, are proved by the results of a<br />

number of projects on natural resources evaluation. Natural resources provide stable flow of<br />

economic revenue. Along with the above mentioned, natural resources are of great social<br />

importance in rural areas, as they provide countrymen living. Natural forest resources are the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Tsibulnikova 2010. Economic estimations in using of the l<strong>and</strong>scapes planning<br />

324<br />

only source of revenue, <strong>and</strong> therefore the only means of subsistence for the majority of<br />

countrymen.<br />

Social <strong>and</strong> ecological significance of natural resources is often underestimated at managerial<br />

decision making due to the absence of any relevant information regarding their full economic<br />

value at the stage of economic analysis (Tsibulnikova, 2003).<br />

2. Methodology<br />

Methodological approaches to natural resources evaluation recommended by UN analysis<br />

department allow to define full economic value of natural objects.<br />

The peculiarity of the methods is that natural resources value is defined as capitalized annual<br />

rent over the period of their full utilization. Estimation of total volume of utilized natural<br />

resources is the key point of this methodology. Nonmarket <strong>and</strong> subjective evaluation methods<br />

allow to define the full volume of utilized resources, including those used by private households,<br />

<strong>and</strong> make overall economic evaluation of territories.<br />

The methods were first approved in Tomsk region for the evaluation of the area between Ob’<br />

<strong>and</strong> Tom’ rivers, with the aim to study the problem of social <strong>and</strong> ecological factors synthesis in<br />

economic evaluation of natural resources capital, <strong>and</strong> develop mechanisms for unique territories<br />

preservation (Adam, Tsibulnikova, 2001).<br />

2.1 Local methodology approbation<br />

To study the problem of combination social <strong>and</strong> ecological factors in the economic estimation<br />

of the territory natural capital the local territory between the rivers Ob <strong>and</strong> Tom'- the unique<br />

natural complex providing a basic need of Tomsk in recreational resources <strong>and</strong> water has been<br />

used. The territory square is 3600 square kilometers, population – 32000 people.<br />

Now there is a real threat of degradation of the territory ecosystem because of amplifying<br />

anthropogenic loading. The natural resources of this territory are a state property, except the<br />

l<strong>and</strong> where there are settlements. About 50 % of the territory is completely limited for any<br />

economical activities.<br />

<strong>Forest</strong> non-wood recourses estimation was based on the survey data that were used to define the<br />

forest products consumption volume <strong>and</strong> Ob-Tomsk inter-river region <strong>and</strong> Tomsk population<br />

expenses to collect <strong>and</strong> deliver the forest products.<br />

There is an opinion, that if the household collecting the wild grow resources for personal needs<br />

it has the income the same as on the market. The results of the research have shown the stable<br />

level of the economical value of the Ob-Tomsk inter-river territory for the local population <strong>and</strong><br />

for the Tomsk city population. That flow is in several times bigger than the wood resources cost.<br />

Also the wildlife resources have the social importance because they provide the major income<br />

for the family budget of the not too wealthy people.<br />

Indirect cost of the Ob-Tomsk inter-river territory using (indirect cost of the forest) is based on<br />

the capacity of the trees to absorb carbon. The calculation was based on the World Bank<br />

methodology.<br />

That’s why in the natural capital estimation of the Ob-Tomsk inter-river territory the major part<br />

of the useful information was received on the base of the questioning. According to the<br />

questioning we could estimate the wood consumption <strong>and</strong> the non wood resources of the forest<br />

<strong>and</strong> also receive information how the local population is ready to bear expenses to preserve this<br />

territory.<br />

Estimation of the animals <strong>and</strong> fish was received on the actual shooting of the animals <strong>and</strong> catch<br />

of the fish data (on the base of the questioning), the market prices, <strong>and</strong> the time spent on the<br />

fishing <strong>and</strong> hunting.<br />

Direct use value of the wood is 125 000 dollars – 3% from the total cost of the forest. The<br />

economical value of the forest is in 4,5 times higher in comparison with the simple wood,<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Tsibulnikova 2010. Economic estimations in using of the l<strong>and</strong>scapes planning<br />

325<br />

because the forest is using by the households as the source of the food products. Applying the<br />

specific estimation methods allow to increase in 30 times the economical value of the forest.<br />

The net value of the Ob-Tomsk inter-river territory forest resources on the base of the survey is<br />

about 2,9 million dollars per year. If the discounting rate will be 3 % <strong>and</strong> the usage time of the<br />

non-wood resources of the forest will be 100 years, the total cost of them is about 91,6 million<br />

dollars. To compare, the total revenue from the complete felling of the Ob-Tomsk inter-river<br />

trees plus the second trees felling 100 years later (the forest renewal period), with the same<br />

discounting rate 3 %, will be 26,9 million dollars.<br />

Comparison ecology services distribution structure shows that practically on all the basic<br />

ecosystem services export to a city essentially exceeds internally consumption. From the<br />

positions of the natural capital movement analysis <strong>and</strong> fair distribution of the natural rent,<br />

directions <strong>and</strong> methods of preservation strategy <strong>and</strong> development of a natural l<strong>and</strong>scape have<br />

been defined. Monetary estimations of territory natural resources have been applied by us as<br />

territorial indicators <strong>and</strong> indicators of steady wildlife management in particular economicgeographical<br />

conditions. (see Figure 1)<br />

After reception of the results the questions of incomes increasing because of the natural capital<br />

<strong>and</strong> creation of conditions for forest products collateral preparation, fishing, cultivations of wild<br />

animals for the hunting tourism became priority.<br />

The received results have allowed to develop actions for preservation <strong>and</strong> development of<br />

territory adjoining to a city because of the reinvestment the natural rent in preservation of a<br />

natural complex, according to sustainable development principles. On the basis of the received<br />

results of that work, an estimation of the area natural resources <strong>and</strong> creation of system of the<br />

ecological-economic account at regional level has been continued.<br />

Figure 1: Economical estimation of the forest natural resources with taking into account<br />

ecological <strong>and</strong> social factor . (Tsibulnikova 2001).<br />

2.2 Development of a system for ecological <strong>and</strong> economical registration in the region<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Tsibulnikova 2010. Economic estimations in using of the l<strong>and</strong>scapes planning<br />

326<br />

Results of the approbation of pecuniary valuation of natural resources in Tomsk region were<br />

spread over the region with the aim to integrate the system of ecological <strong>and</strong> economical<br />

registration into management practice.<br />

Further investigations in the region have shown that fuel-energy resources are most significant<br />

in the structure of natural capital of Tomsk region, comprising up to 96% of its total value.<br />

These factors defined the strategy of Tomsk region development aimed at priority development<br />

of oil <strong>and</strong> gas industry (Adam, Tsibulnikova, 2009).<br />

At the same time, in spite of growing dependence of regional economy on petroleum production,<br />

biological resources significantly account in the structure of natural capital. (see Figure 2).<br />

Such conditions set a task to support stable flow of natural resources <strong>and</strong> ecosystem services of<br />

forest l<strong>and</strong>scapes.<br />

The analysis of information coming from regions has shown that petroleum production does not<br />

play the key role in regional population employment. Petroleum production is realized in three<br />

regions with total population comprising only 9% of Tomsk region population. 25% of the<br />

population of Tomsk region reside in 12 region occupying 50% of the territory of Tomsk region,<br />

where people earn their living mainly by biological resources. Tomsk region is subdivided in 16<br />

municipal regions. Small scale entrepreneurships <strong>and</strong> businessmen are engaged in fishing or<br />

forest resources storage.<br />

In addition to forest foodstuff storage, some regions organize storage of some crude drugs, food<br />

plants, <strong>and</strong> other resources, used by the population in the development of local production. Total<br />

economic value of nonwood forest resources amounts to $979.3 million. The estimation is not<br />

final, as it is made on the basis of 12 regions out of 16. Moreover 8 of them carry out<br />

registration of foodstuff forest resources only, <strong>and</strong> only 2 regions carry out registration of all<br />

kinds of nonwood forest resources. According to peer review, only 30% of total economic value<br />

of forest l<strong>and</strong>scapes is being registered presently.<br />

Economic evaluation of forest resources has proved mushrooms, berries, <strong>and</strong> medicinal herbs to<br />

be the full source of revenue for the local population. Thus, economic value of nonwood forest<br />

resources is 6 times higher than that of wood. At that, mushrooms <strong>and</strong> pine nuts represent the<br />

greatest part in total value (47% <strong>and</strong> 24% of total value accordingly). Due to the absence of<br />

official resources registration system in the state, planning of efficient use <strong>and</strong> preservation of<br />

natural l<strong>and</strong>scapes involve difficulties. As a result, in the course of industrial wood harvest <strong>and</strong><br />

mining, the population loses nonwood resources playing significant role in life support.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Tsibulnikova 2010. Economic estimations in using of the l<strong>and</strong>scapes planning<br />

327<br />

Figure 2: Natural capital structure of the Tomsk Region (economical value of the natural resources)<br />

(Tsibulnikova 2009)<br />

3. Result<br />

Current investigations change approaches to nature management. Economical information is an<br />

instrument providing calculations necessary for substantiation managerial decisions. Tomsk<br />

region administration directed official letters stating the necessity to organize better registration<br />

of profit gained from forests to municipal regions.<br />

Two municipal regions have rather well organized registration procedures providing relevant<br />

information, <strong>and</strong> give preference to forest preservation rather than wood storage. One region<br />

decided to direct funds for berry plantations quality improvement. Another region decided to<br />

create natural recreation zone.<br />

The necessity to preserve natural l<strong>and</strong>scapes often conflict with the aims of social <strong>and</strong> economic<br />

development of regions. Local authorities empowered to manage natural resources, often fail to<br />

take into account that life of many people is dependent on natural resources (Tsibulnikova<br />

2003).<br />

Working out <strong>and</strong> adoption of Tomsk regional statute obliging mining companies to compensate<br />

for losses of natural resources have become the practical application of natural resources<br />

evaluation methods taking into account both social <strong>and</strong> ecological factors. The amount of<br />

compensation is calculated from the cost of natural resources directly or indirectly involved in<br />

the process of mining.<br />

4. Discussion<br />

1. Application of pecuniary valuation in natural resources evaluation methodology with the<br />

frame of l<strong>and</strong>scape <strong>and</strong> ecological investigations allows to consider full economic value of<br />

nature at managerial decision making.<br />

2. Economic evaluation taking into account social <strong>and</strong> ecological factors allow to define real<br />

material <strong>and</strong> money flows in nature management.<br />

3. Pecuniary valuation of natural resources taking into account social factor show economic <strong>and</strong><br />

geographical peculiarities of the territory, indicate nature management stability, <strong>and</strong> can be used<br />

both for natural <strong>and</strong> anthropogenic l<strong>and</strong>scape management.<br />

4. Economic <strong>and</strong> geographical analysis of nature management in Tomsk region has shown that<br />

in the effort to provide stable development of Tomsk region, preservation of natural ability of<br />

forest to provide foodstuff, crude drug, <strong>and</strong> other forest resources, will contribute to better<br />

economic efficiency<br />

5. The analysis of money flows coming from nonwood forest resources gives the ground for<br />

local authorities to make right choice in territories use <strong>and</strong> define such forest resources as<br />

recreation, creation <strong>and</strong> further utilization of forest plantations, growing of forest fruit, berry,<br />

ornamental plants, <strong>and</strong> medicinal herbs, as priority ones.<br />

6. In the effort to provide stable development of Tomsk region, preservation of natural ability of<br />

forest to provide foodstuff, crude drug, secondary forest resources, <strong>and</strong> accessory products of<br />

forest exploitation without l<strong>and</strong>scape deterioration, will contribute to better economic efficiency.<br />

Such course can provide multiplicative effect from animal <strong>and</strong> fish natural habitat preservation,<br />

<strong>and</strong> forest recreation development for amateur hunting <strong>and</strong> fishing.<br />

7. Evaluation of money flows in nature management allows not only to plan arrangements for<br />

natural resources preservation, but also to define volumes <strong>and</strong> mechanisms for compensation in<br />

case of l<strong>and</strong>scape deterioration.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Tsibulnikova 2010. Economic estimations in using of the l<strong>and</strong>scapes planning<br />

328<br />

References<br />

Adam A. M., Tsibulnikova M. R., Laptev N. I. Regional ecological policy: Tomsk experience. –<br />

Moscow: Institute for Stable Development/Centre for Ecological Policy of Russia, 2009.-<br />

60 p.<br />

Adam A. M., Tsibulnikova M. R. Application of pecuniary valuation of natural resources as the<br />

indicator of stable nature management. Selected Parers presented at 7 th International<br />

Scientific <strong>and</strong> Practical Conference «Natural <strong>and</strong> intellectual resources of Siberia<br />

(Sibresurs-7-2001)», Tomsk, 2001. P. 154-158.<br />

Economical basis for prevention of disputes in nature management by the example of the area<br />

between Ob’ <strong>and</strong> Tom’ rivers. Scientific paper № 7/2000. Yaroslavl. 2000. 107 p.<br />

Tsibulnikova M. R. The usage of natural resources money estimations in control of nature<br />

management on Ob-Tom interriver. Environment of Siberia, the Far East, <strong>and</strong> the<br />

Arctic/Selected Papers presented at the International Conference, Tomsk, 2001. P. 422-<br />

426.<br />

Tsibulnikova M. R. Natural resources evaluation in social <strong>and</strong> economical development of<br />

regions. Issues of utilization <strong>and</strong> preservation of natural resources of Central Siberia.<br />

Issue 5. Krasnoyarsk: KNIIG&MS, 2003. P. 12-13.<br />

Tsibulnikova M. R. Economical aspects of nature management in regional development/Issues<br />

of geology <strong>and</strong> geography of Siberia. Bulletin of Tomsk State University № 3, TSU,<br />

Tomsk, 2003. P. 69-71<br />

Adam A. M., Tsibulnikova M. R. Use of biological resources as one of the directions for the<br />

increase in Tomsk region development stability. «Natural <strong>and</strong> intellectual resources of<br />

Siberia (Sibresurs-15-2009)». Slected Papers presented at 15 th International Scientific <strong>and</strong><br />

Practical Conference, Irkutsk, October 5-7, 2009./responsible editor V. N. Maslennikov. –<br />

Tomsk: SAN VSh; V-Spektr, 2009. 304 p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Section 5<br />

Monitoring l<strong>and</strong>scape change


M. Akbarzadeh & E. Kouhgardi 2010. Mapping <strong>and</strong> monitoring l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> –use changing using RS <strong>and</strong> GIS<br />

331<br />

Mapping <strong>and</strong> Monitoring l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> –use changing using RS<br />

<strong>and</strong> GIS. Case study: Kaleybar, Iran<br />

M. Akbarzadeh 1* & E. Kouhgardi 2<br />

1 Islamic Azad University, Myianeh Branch, East Azarbaijan, Iran.<br />

2 Islamic Azad University, Boushehr Branch, Boushehr, Iran.<br />

Abstract<br />

Arasbaran is located in East Northern part of East Azerbijan province. Because of the having<br />

montainus forests <strong>and</strong> special st<strong>and</strong>ing condition, fragile ecosystem is created.<br />

In this study, maximum likelihood supervised classification <strong>and</strong> post – classification change<br />

detection techniques were applied to l<strong>and</strong> sat images acquired in 1987 <strong>and</strong> 2002, to map l<strong>and</strong><br />

cover changes in the North Western forests of Iran.<br />

A supervised classification was carried out on the six reflective b<strong>and</strong>s for the two images<br />

individually.<br />

Key words: Iran, Arasbaran, Remote sensing, GIS, L<strong>and</strong> cover<br />

1. Introduction<br />

Remote sensing techniques have recently recived lots of attentions in agriculture <strong>and</strong> natural<br />

resources. Natural resources <strong>and</strong> environmental conservation need a lot of attention especially<br />

on under devloped countries. Using remote sensing techniques <strong>and</strong> sateliate data for evaluation<br />

of environmental changes is rapidly growing.<br />

2. Methodology<br />

2.1 Study area<br />

The study area is located in North West of IRAN <strong>and</strong> the North Eastern of the East Azerbijan<br />

province, which called Arasbaran, is a mountainous area with elevation between 300 <strong>and</strong> 2700<br />

meters above the see level <strong>and</strong> very near to the Caspian Sea.<br />

o<br />

o<br />

o<br />

o<br />

The area is located between 38 .48′<br />

<strong>and</strong> 39 .01′<br />

latitude <strong>and</strong> between 47 .14′<br />

<strong>and</strong> 46 .51′<br />

longitude. It covers diversity of elevation, slope, population <strong>and</strong> l<strong>and</strong> use <strong>and</strong> includes a variety<br />

of seashore rivers, etc. There are above 785 plant species in this forest that 97 species of them<br />

are woody (Sagheb et al, 2004).<br />

Arasbaran includes 11 basins. Kaleybar Chay is one of them, which is our study area. Aras river<br />

is in the Northern boundary of the study area.<br />

2.2 Materials<br />

Three sets of material were used here.<br />

* Corresponding author. Tel.: +98914 406 6696 - Fax: +98 423 22 400 85<br />

Email address: m.akbarzadeh@m_iau.ac.ir<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Akbarzadeh & E. Kouhgardi 2010. Mapping <strong>and</strong> monitoring l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> –use changing using RS <strong>and</strong> GIS<br />

332<br />

First, l<strong>and</strong> sat Thematic Mapper (TM) <strong>and</strong> Enhanced Thematic Mapper (ETM+) images<br />

acquired on 20 July 2002 <strong>and</strong> 19 July 1987, respectively. L<strong>and</strong> sat 4 <strong>and</strong> 5 carry the TM sensors.<br />

Second, digital topographic maps digitized from hard copy topographic maps with scale of<br />

1:50,000 were made of IRANIAN surveying center <strong>and</strong> used mainly for geometric correction of<br />

the satellite data <strong>and</strong> for some ground truth information.<br />

Finally, ground information was collected between 1997 until 2002for the purpose of supervised<br />

classification <strong>and</strong> classification accuracy assessment.<br />

2.3 Geometric correction<br />

Accurate per- pixel registration of multi- temporal remote sensing data is essential for change<br />

detection since the potential exists for registration errors to be interpreted as l<strong>and</strong> cover <strong>and</strong><br />

l<strong>and</strong>- use change, leading to an overestimation of actual <strong>Change</strong>(stow, 1999).<br />

<strong>Change</strong> detection analysis is performed on a pixel – by- pixel basis, therefore a misregistration<br />

greater than one pixel will provide an anomalous result of that pixel. To overcome this problem,<br />

the root mean – square error (RMSE) between any two dates should not exceed 0.5 pixels (Lune<br />

tta& Elvidge 1998).<br />

In this study geometric correction was carried out using ground control points from topographic<br />

maps with scale of 1:5000 produced in 1986 by Iranian Army Surveying Center (IASC).To<br />

geocode the image of 2002, then this image was used to register the image of 1987.<br />

The RMSE between the two images was less than 0.4 pixel which is acceptable. The RMES<br />

could be defined as the deviations between GCP <strong>and</strong> GP locations as predicted by the fitted –<br />

polynomial <strong>and</strong> their actual locations.<br />

The goal of image enhancement is to improve the visual interpretability of an image by<br />

increasing the apparent distinction between the features.<br />

The process of visually interpreting digitally enhanced imagery attempts to optimize the<br />

complementary abilities of the human mind <strong>and</strong> the computer.The mind is excellent at<br />

interpreting spatial attributes on an image <strong>and</strong> is capable of identifying obscure or subtle<br />

features (Lilles<strong>and</strong> & Kiefer, 1994).<br />

Contrast stretching was applied on the two images <strong>and</strong> two false color composites (FCC) were<br />

produced.<br />

These FCC were visually interpreted using on screen digitizing in order to delineate l<strong>and</strong> cover<br />

classes that could be easily interpreted such as village.<br />

Some classes were spectrally confused <strong>and</strong> could not be separated well by supervised<br />

classification <strong>and</strong> hence visual interpretation was required to separate them.<br />

2.4 Image classification<br />

L<strong>and</strong> cover classes are typically mapped from digital remotely sensed data through the process<br />

of a supervised digital image classification (Campbell, 1987), (Zobeyri & Daleki, 1988). The<br />

over all objective of the image classification procedure is to automatically categorize all pixels<br />

in an image <strong>and</strong> to l<strong>and</strong> cover classes or themes (Lilles<strong>and</strong> & Kiefer, 1994).<br />

The maximum likelihood classifier quantitatively evaluates both the variance <strong>and</strong> covariance of<br />

the category spectral response patterns when classifying an unknown pixel so that it is<br />

considered to be one of the most accurate classifiers since it is based on statistical parameters.<br />

Supervised classification was done using ground chek points <strong>and</strong> digital topographic maps of<br />

the study area.<br />

The area was classified into seven main classes: High density forest, low density forest,<br />

rangl<strong>and</strong>s, agriculture, bare l<strong>and</strong>, river, <strong>and</strong> village.<br />

Description of theses l<strong>and</strong> cover classes are presented in Table 1.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Akbarzadeh & E. Kouhgardi 2010. Mapping <strong>and</strong> monitoring l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> –use changing using RS <strong>and</strong> GIS<br />

333<br />

In order to increase the accuracy of l<strong>and</strong> cover mapping of the two images, a cillary data <strong>and</strong> the<br />

result of visual interpretation were integrated with the classification result using GIS in order to<br />

improve the classification accuracy of the classified image.<br />

Table1. Classification details of l<strong>and</strong> covers classes in study area<br />

Class<br />

Details<br />

<strong>Forest</strong> st<strong>and</strong>s<br />

Range l<strong>and</strong><br />

Cropl<strong>and</strong><br />

Bare l<strong>and</strong><br />

River<br />

Village<br />

Mountainous forest st<strong>and</strong>s ,which almost coppice<br />

st<strong>and</strong>s <strong>and</strong> low<br />

forests including in study area. The area have the biggest forest st<strong>and</strong> in North<br />

West of Iran.<br />

Areas which no ability for agriculture <strong>and</strong> forestry<br />

grasses. Some range l<strong>and</strong>s is part of<br />

positive line of ecosystem changing.<br />

but have<br />

changing biocenose at<br />

L<strong>and</strong>s which cultivated vegetables, fruits, <strong>and</strong> annual crops. These crops are<br />

irrigated mainly from the water of rivers Aras Ilghine chay , or rain water <strong>and</strong> or<br />

ground water.<br />

L<strong>and</strong> areas of exposed soil surface as influenced by human impact mainly or<br />

natural causes. That is the last trying of ecosystem to protect its biocenose.<br />

Aras <strong>and</strong> Ilghineh chay are two main rivers in study area <strong>and</strong> main resources for<br />

irrigation of cropl<strong>and</strong>s <strong>and</strong> human uses, after this, the area including springs.<br />

The biggest population community where human society is in it, at the study area.<br />

The place of villages did not change during the time of this study so they were<br />

not included in our classifications.<br />

2.5 L<strong>and</strong> cover/use change detection<br />

Regardless of the techniques used, the success of change detection from imagery will depend<br />

on both the nature of the change involved <strong>and</strong> the success of the image pre processing <strong>and</strong><br />

classification procedures. If the nature of change within a particular scene is either abrupt or at a<br />

scale appropriate to the imagery collected then change should be relatively easy to detect.<br />

Problems occur only if spatial changes are subtly distributed <strong>and</strong> hence not obvious within any<br />

image pixel (Milne, 1988, Darvishsefat et al, 2002).<br />

In the case of the study area chosen, field observation <strong>and</strong> measurement have shown that the<br />

change between the image collection dates was both marked <strong>and</strong> abrupt .In this study post -<br />

classification change detection technique was applied. Post classification is the most abvious<br />

method of change detection, which requires the comparison of independently produced<br />

classified images. The CROSSTAB module of Ilwis software was used for performing cross<br />

tabulation analysis. CROSSTAB performs two operations.<br />

The first image cross tabulation in which the categories of one image are compared with those<br />

of a second image <strong>and</strong> of the number of cells in each combination is kept in tabulation. The<br />

result of this operation is a table listing the tabulation totals as well as several measures of<br />

association between the images.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Akbarzadeh & E. Kouhgardi 2010. Mapping <strong>and</strong> monitoring l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> –use changing using RS <strong>and</strong> GIS<br />

334<br />

The second operation that CROSSTAB offers is cross- classification. Cross- classification can<br />

be described as a multiple overly showing all combinations of the logical AND operation<br />

(Shalaby.A&Tateishi.R, 2007).<br />

The result is a new image that shows the locations of all combinations of the categories in the<br />

orginal images<br />

A legend is automatically produced showing these combinations' cross- classification thus<br />

produce a map representation of all non- zero entries in the cross- tabulation table.<br />

3. Result<br />

The color composites generated from b<strong>and</strong>s 4, 3 <strong>and</strong> were visually interpreted through on screen<br />

digitizing. The visual interpretation gave a general idea about the forms of l<strong>and</strong> cover changes<br />

over the period. Many farml<strong>and</strong>s <strong>and</strong> new roads <strong>and</strong> range l<strong>and</strong>s were noticed in the image of<br />

2002 <strong>and</strong> not in the one of 1987. A noticeable change is detected in areas of forest st<strong>and</strong>s, lime<br />

stone <strong>and</strong> shale Quarries in the Northern <strong>and</strong> South eastern parts of the study area where part of<br />

the forest st<strong>and</strong>s, range l<strong>and</strong>s <strong>and</strong> cropl<strong>and</strong> were converted to range l<strong>and</strong>, cropl<strong>and</strong> <strong>and</strong> bare<br />

l<strong>and</strong>s. On screen digitizing was carried out for forest, range l<strong>and</strong>s, cropl<strong>and</strong>, bare l<strong>and</strong>, river,<br />

village classes as well as the road network. Supervised classification using all reflective b<strong>and</strong>s<br />

of two images acquired on 20 July 2002 <strong>and</strong> 19 July 1987 was carried out using maximum likeli<br />

hood classifier. In order to increase the accuracy of l<strong>and</strong> cover mapping of the two images,<br />

ancillary data <strong>and</strong> the result of visual interpretation were integrated with the classification<br />

results using GIS. The module used the overlay module in Ilwis 3.0 Academic Software.<br />

Through the overlay proccess, areas which were misclassified in the village forest, range l<strong>and</strong>s<br />

<strong>and</strong> cropl<strong>and</strong> <strong>and</strong> river classes were relabeled to the correct classes using the layer of visual<br />

interpretation.<br />

This overlying of the visual interpretation on the result of the classification led to the increase in<br />

the overall accuracies by about 10 percent for both images.<br />

The lowest accuracy was for range l<strong>and</strong> <strong>and</strong> forest that could be explained by the fact that the<br />

study area is semi – humid <strong>and</strong> the vegetation intensity is between range l<strong>and</strong>s <strong>and</strong> forest st<strong>and</strong>s,<br />

especially in ecotons which gradually led to the confusion with them.<br />

Remote sensing data <strong>and</strong> GIS provide opportunities for integrated analysis of spatial data.<br />

Cross- tabulation performs image Cross- tabulation in which the categories of one image are<br />

compared with those of a second image <strong>and</strong> tabulation is kept of the number of cells in each<br />

combination.<br />

Post - classification change detection technique was carried out, through Cross- tabulation GIS<br />

module for the classification results of 1987 <strong>and</strong> 2002 images in order to produce change image<br />

<strong>and</strong> statistical data about the spatial distribution of different l<strong>and</strong> cover changes <strong>and</strong> nonchange<br />

areas .<br />

We have to take in to consideration the accuracy of the classification of different classes since<br />

the error of the classification will affect the accuracy of the change detection figures.<br />

L<strong>and</strong> degradation processes in the study area are: degradation of range l<strong>and</strong> due to overgrazing<br />

<strong>and</strong> the converting of range l<strong>and</strong> to cropl<strong>and</strong> due to low incoming <strong>and</strong> bad economical condition<br />

of peopls who living in study area.<br />

Mismanagement of natural resources specialy forest st<strong>and</strong>s is another cause of forest<br />

degradation. In this view point people who living in villages need fuel <strong>and</strong> means <strong>and</strong> materials<br />

for their lifes. So with mismanagement of natural resources, the forest st<strong>and</strong>s is the easy <strong>and</strong><br />

best answer for solving their problems.<br />

This could be seen on the l<strong>and</strong> cover / l<strong>and</strong> use map of 2002 where the crop l<strong>and</strong> areas have<br />

increased from 400.52 ha crop l<strong>and</strong> in 1987 to 654.76 ha crop l<strong>and</strong> in 2002.( For more see<br />

table2).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Akbarzadeh & E. Kouhgardi 2010. Mapping <strong>and</strong> monitoring l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> –use changing using RS <strong>and</strong> GIS<br />

335<br />

Table2. <strong>Change</strong> detection details of study area<br />

Class Name 1987 (area in ha) 2002(area in ha)<br />

Crop l<strong>and</strong><br />

Bairsoli<br />

H.<strong>Forest</strong><br />

L.<strong>Forest</strong><br />

Rangel<strong>and</strong><br />

River<br />

Total area<br />

400.52<br />

21128.71<br />

5089.26<br />

14036.59<br />

9049.88<br />

303.7<br />

50008.66<br />

654.76<br />

21931.34<br />

4086.3<br />

13987.45<br />

9021.29<br />

327.52<br />

50008.66<br />

The next l<strong>and</strong> degradation process cause in the study area is water erosion <strong>and</strong> wind erosion,<br />

which are so power full, in l<strong>and</strong>s without vegetation cover or low vegetation cover.<br />

Water <strong>and</strong> wind erosion let to the removal of the relatively fertile top soil <strong>and</strong> this could lead to<br />

desertification.<br />

4. Discussion<br />

The objective of this study was to provide a recent prespective for l<strong>and</strong> cover types <strong>and</strong> l<strong>and</strong><br />

cover changes that have taken place in the last fourteen years, to integrate visual interpretation<br />

with supervised classification using GIS <strong>and</strong> to examine the capabilities of integrating remote<br />

sensing <strong>and</strong> GIS in studying the spatial distribution of different l<strong>and</strong> cover changes.<br />

The area of forest st<strong>and</strong>s has decreased considerably. Integrating GIS <strong>and</strong> remote sensing<br />

provided valuable information on the nature of l<strong>and</strong> cover changes especially on the area <strong>and</strong><br />

spatial distribution of different l<strong>and</strong> cover changes.<br />

The main causes of l<strong>and</strong> degradation in the study area are convertion of forest st<strong>and</strong>s to range<br />

l<strong>and</strong>s, range l<strong>and</strong>s to crop l<strong>and</strong>s, <strong>and</strong> convertion of crop l<strong>and</strong>s to bare l<strong>and</strong>s. This problem show<br />

sere to be continued for disclimax <strong>and</strong> needs to be seriously studied, through<br />

References<br />

Akbarzadeh, M., 2007. Using RS <strong>and</strong> GIS for L<strong>and</strong> evaluation in Arasbaran Iran. Ph.D. thesis,<br />

IAU, Science <strong>and</strong> Research University, Tehran, Iran.<br />

Darvish sefat, A., Makhdum, M., <strong>and</strong> Jafarzadeh, H., 2002. Using GIS <strong>and</strong> RS in natural<br />

resources.Tehran University Press, 250 p.<br />

Milne, A., 1988. <strong>Change</strong> detection analysis using L<strong>and</strong>sat imagery a review of methodology.<br />

Proceedings of IGARSS, 88 symposium, 541–544.<br />

Stow, D., 1999. Reducing mis-registration effects for pixel-level analysis of l<strong>and</strong>-cover change.<br />

International Journal of Remote Sensing, 20: 2477–2483.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Baran-Zgłobicka & W. Zgłobicki 2010. <strong>Forest</strong> patches in agricultural l<strong>and</strong>scapes (loess areas of SE Pol<strong>and</strong>)<br />

336<br />

<strong>Forest</strong> patches in agricultural l<strong>and</strong>scapes<br />

(loess areas of SE Pol<strong>and</strong>)<br />

Bogusława Baran-Zgłobicka & Wojciech Zgłobicki<br />

Department of Geology <strong>and</strong> Lithosphere Conservation,<br />

Maria Curie - Sklodowska University, Pol<strong>and</strong><br />

Abstract<br />

A mosaic of naturally <strong>and</strong> anthropogenically conditioned patches of terrain cover, which are<br />

diversified as regards the kind, size <strong>and</strong> shape, appears in the agricultural l<strong>and</strong>scapes. The<br />

research was carried out in the 2 test areas located in SE Pol<strong>and</strong>. They cover an area of 28 km2<br />

<strong>and</strong> 35 km2. This is the area of the appearance of a specific l<strong>and</strong>scape in the structure of which,<br />

a group of forms characteristic for its loess relief is dominant. The analysis of the changes in the<br />

l<strong>and</strong> use were carried out on the basis of the three maps comparison four maps, representing the<br />

years 1840, 1890, 1935, 1997. The spatial analysis consisted of the assessment of changes of<br />

forest patches areas as well as some chosen statistical parameters prepared for the four time<br />

periods: 1840-1890, 1890-1935, 1935-1997, 1890-1997. The studies revealed differences in the<br />

scope of changes in the percentage of forest cover in the test areas.<br />

Keywords: agricultural l<strong>and</strong>scape, l<strong>and</strong> use changes, loess areas<br />

1. Introduction<br />

Agricultural l<strong>and</strong>scapes are characterised by a mosaic of naturally <strong>and</strong> anthropogenically<br />

conditioned patches of l<strong>and</strong> cover that are diverse in kind, size <strong>and</strong> shape. Human activity is the<br />

main factor influencing the changeability of forms of l<strong>and</strong> use in time <strong>and</strong> space. However, in<br />

many cases the influence of abiotic environment components can also be significant. A good<br />

underst<strong>and</strong>ing of the historical determinants of changes in l<strong>and</strong> usage allows for a better<br />

anticipation of the tendencies developing in a l<strong>and</strong>scape, <strong>and</strong> also enables a more effective<br />

l<strong>and</strong>scape management (Bork 1989, Dotterweich 2008).<br />

The research was carried out in two test areas, Wąwolnica (28 km 2 ) <strong>and</strong> Wilczyce (35 km 2 ),<br />

located in SE Pol<strong>and</strong>. The structure of the unique l<strong>and</strong>scape occurring in this area is dominated<br />

by a group of forms characteristic of loess relief. The deep Bystra <strong>and</strong> Opatówka river valleys<br />

form the morphological axes of the test areas. The plateaus <strong>and</strong> slopes are dissected by a system<br />

of dry valleys <strong>and</strong> gullies. The Wąwolnica area exhibits a greater diversity of relief as steep<br />

slopes <strong>and</strong> gullies occupy a larger part of the l<strong>and</strong>scape (see Table 1). These areas were<br />

cultivated as early as the Neolithic, while the next strong expansion of settlements <strong>and</strong> farming<br />

began in the mediaeval period (Nogaj-Chachaj 2004). Nowadays, arable l<strong>and</strong>s predominate in<br />

the l<strong>and</strong> use structure, occupying from 55% to 63% of the whole area. The percentage of forest<br />

cover is quite high in the Wawolnica test area (18%), while it is significantly lower in Wilczyce<br />

(9%). It is a result of a large proportion of areas where cultivation has been discontinued<br />

because most of them are occupied by gullies <strong>and</strong> steep slopes (Baran-Zgłobicka <strong>and</strong> Zgłobicki<br />

2004).<br />

A typical community growing in gullies <strong>and</strong> on steep slopes nowadays are Tilio-Carpinetum<br />

forests; tree st<strong>and</strong>s are composed of hornbeam (Carpinus betulus) with an admixture of smallleaved<br />

lime (Tilia cordata), Norway maple (Acer platanoides) <strong>and</strong> pedunculate oak (Quercus<br />

robur). Protected species occur among herbal plants (Kucharczyk 1992).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Baran-Zgłobicka & W. Zgłobicki 2010. <strong>Forest</strong> patches in agricultural l<strong>and</strong>scapes (loess areas of SE Pol<strong>and</strong>)<br />

337<br />

2. Methodology<br />

The analysis of the changes in l<strong>and</strong> use was carried out by comparing three maps: Karte des<br />

westlichen Rußl<strong>and</strong>s 1: 100 000 (representing the year 1890), Tactical map of Pol<strong>and</strong> WIG 1:<br />

100 000 (representing the year 1935), <strong>and</strong> a modern l<strong>and</strong> use map 1: 25 000 (representing the<br />

year 1997, elaborated based on aerial photographs <strong>and</strong> field mapping). The first two maps were<br />

digitized <strong>and</strong> calibrated. The next step consisted of the creation of maps (layers) of wooded<br />

areas with the use of ArcView software. The map compilation <strong>and</strong> the selection of chosen<br />

separations enabled the analysis of how the percentage of forest cover changed over time as<br />

well as the examination of the links between those processes <strong>and</strong> natural determinants. A map<br />

dating back to 1840 was also available but did not lend itself to cartometric analysis. Hence,<br />

only an approximate percentage was calculated for the forest cover in that year.<br />

The above-mentioned cartographic materials indicate that woodl<strong>and</strong>s <strong>and</strong> forests are the only<br />

l<strong>and</strong> use category that could be examined in relation to such a long period of time. At the same<br />

time, they constitute the most “natural” component of the l<strong>and</strong> cover. The spatial analysis<br />

included the assessment of changes in the area of forest patches as well as selected statistical<br />

parameters prepared for three periods: 1890-1935, 1935-1997, 1890-1997. In addition, an<br />

examination has been conducted with respect to the character of the abiotic components in areas<br />

covered by forests in various periods as well as in areas where forestation or deforestation took<br />

place.<br />

3. Results<br />

The studies revealed differences in the scope of changes in the percentage of forest cover in the<br />

test areas.<br />

- In the Wąwolnica test area, the forest cover decreased by 2.8 km 2 over the last 100 years; the<br />

lowest forest cover occurred in 1935 (11%)<br />

- In the Wilczyce test area, the percentage of forest cover increased more than fivefold over the<br />

last 100 years. At the end of the 19 th <strong>and</strong> beginning of the 20 th century it was less than 0.5% (see<br />

Figure 1)<br />

- In Wilczyce, gullies are the only contemporary form of relief with a considerable forest cover.<br />

In Wąwolnica, the percentage of the forest cover on steep slopes is also relatively high.<br />

- In Wąwolnica, an increased number of forest patches was observed, along with a decreased<br />

mean patch area. The diversity of the shapes of the patches has also increased (see Table 2).<br />

Nowadays, the mean patch area is particularly small in the case of Wilczyce.<br />

- The loess plateaus as well as gentle <strong>and</strong> medium slopes were deforested in the Wąwolnica test<br />

area (see Table 3), while in the Wilczyce test area this process was not observed.<br />

- In both test areas, the forests occupied the area of gullies <strong>and</strong> steep slopes.<br />

4. Discussion<br />

The presented results show the rationality of changes as regards the usage of the agricultural<br />

production space. In Wąwolnica, the plateaus <strong>and</strong> gentle slopes, convenient for agricultural<br />

cultivation, were deforested during the last 100 years, whereas in Wilczyce this process took<br />

place before the 19 th century. After World War II, an increase of forest areas occurred in both<br />

test areas. The succession of forests occurred in areas with a significant threat of soil <strong>and</strong> gully<br />

erosion. Nowadays, this process also occurs in connection with a decline in the profitability of<br />

agricultural production.<br />

The differences in the percentage of forest cover between the test areas should be linked with<br />

natural <strong>and</strong> socioeconomic conditions. The clearly larger area covered by forests in Wąwolnica<br />

results from the relief of the area, i.e. a large proportion of steep slopes <strong>and</strong> gullies. The<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Baran-Zgłobicka & W. Zgłobicki 2010. <strong>Forest</strong> patches in agricultural l<strong>and</strong>scapes (loess areas of SE Pol<strong>and</strong>)<br />

338<br />

Wilczyce area was characterised by a more intensive agricultural activity connected with the<br />

existence of vast ecclesiastical estates in this area.<br />

From the perspective of ecology, the current structure of forest patches is not favourable<br />

because forested areas are small <strong>and</strong> diverse in shape. On the other h<strong>and</strong>, forests in gullies <strong>and</strong><br />

on steep slopes represent the only forest enclaves in the agricultural l<strong>and</strong>scape. Such a situation<br />

clearly indicates the role of geodiversity as a determinant of biodiversity.<br />

In some cases, however, the changes in usage structure were not determined by nature-related<br />

factors. In Wąwolnica, considerable deforestation took place within the medium slopes, whereas<br />

in the case of steep slopes the forestation area represented less than 50% of the previously<br />

deforested areas. At the same time, some changes in the agricultural character of l<strong>and</strong> use are<br />

necessary in some parts of the analysed areas – most importantly in order to limit the threat of<br />

erosion.<br />

References<br />

Baran-Zgłobicka B., Zglobicki W. 2004: Structure of the agricultural loess l<strong>and</strong>scape (on the<br />

example of the western fragment of the Naleczow Plateau). Fizyčna geografija ta<br />

geomorfologija. Mižvidomčyj naukovyj zbirnik, 46: 12-18.<br />

Bork H.-R. 1989. Soil erosion during the past Millennium in Central Europe <strong>and</strong> its significance<br />

within the geomorphodynamics of the Holocene. Catena Supplement, 15: 221-131.<br />

Dotterweich M. 2008. The history of soil erosion <strong>and</strong> fluvial deposits in small catchments of<br />

central Europe: Deciphering the long-term interaction between humans <strong>and</strong> the<br />

environment. Geomorphology, 101: 192-208..<br />

Kucharczyk M. 1992. Roślinność i flora. Kazimierski Park Krajobrazowy. In: Tadeusz Wilgat<br />

(Ed.) System obszarów chronionych województwa lubelskiego. UMCS, TWWP, LFOŚN,<br />

Lublin: 76-81.<br />

Nogaj-Chachaj J. 2004: O roli człowieka w przekształcaniu środowiska przyrodniczego w<br />

holocenie na Płaskowyżu Nałęczowskim. In: Jerzy. Libera <strong>and</strong> Anna Zakościelna (Eds.)<br />

Przez pradzieje i wczesne średniowiecze. Wydawnictwo UMCS, Lublin: 63-72.<br />

Table 1: Relief <strong>and</strong> l<strong>and</strong> use within test areas<br />

Wąwolnica Wilczyce<br />

Forms of relief<br />

Bottoms of valleys 9 18<br />

Gentle slopes (3-6 o ) 20 17<br />

Moderate slopes (6-12 o ) 16 23<br />

Steep slopes (> 12 o ) 9 4<br />

Gullies 6 4<br />

Plateau tops 42 34<br />

L<strong>and</strong> use<br />

Arable l<strong>and</strong>s 55 63<br />

Orchards 3 14<br />

Plantations 5 0.5<br />

Grassl<strong>and</strong>s/pastures 14 7<br />

<strong>Forest</strong>s 18 9<br />

Wastel<strong>and</strong>s 2 2<br />

Built-up areas 4 4.5<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Baran-Zgłobicka & W. Zgłobicki 2010. <strong>Forest</strong> patches in agricultural l<strong>and</strong>scapes (loess areas of SE Pol<strong>and</strong>)<br />

339<br />

Table 2: <strong>Change</strong>s in the structure of forest patches<br />

<strong>Forest</strong> area <strong>Forest</strong> Number Mean patch area Mean shape index<br />

[km 2 ] cover [%] of patches [ha]<br />

Wąwolnica test area<br />

1890 7.9 28 12 64 1.9<br />

1935 3.1 11 33 10 1.4<br />

1997 5.1 18 37 13 2.2<br />

Wilczyce test area<br />

1997 3.1 9 87 3.7 1.7<br />

Table 3: <strong>Change</strong>s in the forest cover within different forms of relief (Wąwolnica test area)<br />

Plateau tops<br />

Gentle<br />

slopes<br />

Medium<br />

slopes<br />

Steep slopes Gullies Bottoms<br />

of valleys<br />

1890 21 27 35 39 61 9<br />

1935 9 10 13 20 33 2<br />

1997 8 13 20 30 86 4<br />

Figure 1: Tendencies of changes in the forest cover in the test areas<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Beldade & T. Panagopoulos 2010. Integrating esthetical <strong>and</strong> ecological values at the central Asia l<strong>and</strong>scape change<br />

340<br />

Integrating esthetical <strong>and</strong> ecological values at the central Asia<br />

l<strong>and</strong>scape change<br />

Joana Beldade & Thomas Panagopoulos<br />

Centre of Spatial Research <strong>and</strong> Organizations (CIEO), University of Algarve, Campus<br />

de Gambelas, 8000 Faro, Portugal<br />

Abstract<br />

Desertification is the gradual transformation of usable l<strong>and</strong> into desert; is usually caused by<br />

climate change or by destructive use of the l<strong>and</strong>. The present study examines how a person<br />

interprets desert l<strong>and</strong>scapes using phenomenological methodology. In ecological aesthetics<br />

pleasure is derived from knowing on how the part of the l<strong>and</strong>scape relate to the whole. The<br />

objective of the present paper is to describe how to integrate l<strong>and</strong>scape aesthetics into l<strong>and</strong>scape<br />

planning of the Central Asia Silk Road area. A questioner was used as research tool to measure<br />

l<strong>and</strong>scape preference. The questions were constructed following the principles of<br />

phenomenology. After qualitative <strong>and</strong> quantitative analysis, 7 main categories of cognitive<br />

aspects of natural l<strong>and</strong>scapes were identified. Each l<strong>and</strong>scape evoked each participant's<br />

memories <strong>and</strong> background, <strong>and</strong> altered through these influences, imagination/association,<br />

impression, aesthetic judgments, <strong>and</strong> meaning <strong>and</strong> attractiveness of nature. All the above played<br />

their role in the evaluation of desert l<strong>and</strong>scapes.<br />

Keywords: Desert l<strong>and</strong>scape, aesthetics, l<strong>and</strong>scape assessment, phenomenology<br />

1. Introduction<br />

There is mounting evidence pointing to the relationship between climate change effects <strong>and</strong><br />

l<strong>and</strong>scape preferences (Cambers, 2009). L<strong>and</strong>scape preferences <strong>and</strong> perceptions usually are<br />

influenced by demographic elements such as age, gender <strong>and</strong> ethnicity; as well as culture<br />

(Brunson & Shelby, 1992). With few exceptions, most of those studies have examined<br />

perceptions of environments within relative non-dynamic periods. There is therefore a scarcity<br />

of studies on perceptions of desert l<strong>and</strong>scapes that are undergoing reclamation. In light of<br />

climate change, such investigations are crucial to the underst<strong>and</strong>ing of adaptation processes. It is<br />

argued, that to a great extent, tourists’ perceptions have been left out of the debate on desert<br />

l<strong>and</strong>scape adaptation <strong>and</strong> climate change.<br />

Given that desertification within major tourism destinations is an inevitable consequence of<br />

climate change a deeper underst<strong>and</strong>ing of user perceptions can provide a holistic view to<br />

adaptation processes. As the morphological structure of desert l<strong>and</strong>scapes changes, social<br />

driving forces will have to change not only in the way in which they visualize these areas but<br />

also in how they relate to these natural environments <strong>and</strong> how sense of place changes in those<br />

spaces.<br />

The idea of beauty in l<strong>and</strong>scapes has changed during the history of civilization <strong>and</strong> aesthetics<br />

has been a topic of debate for philosophers, artists <strong>and</strong> architects since at least the time of<br />

Socrates (Thorn <strong>and</strong> Huang, 1991; Carlson, 2002). At present, aesthetics is being taken into<br />

consideration by environmental managers <strong>and</strong> policy makers (Canter, 1996). Symmetry <strong>and</strong><br />

other classical rules such as the ‘golden mean’ were the most important components of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Beldade & T. Panagopoulos 2010. Integrating esthetical <strong>and</strong> ecological values at the central Asia l<strong>and</strong>scape change<br />

341<br />

l<strong>and</strong>scape beauty in Greek <strong>and</strong> Roman gardens two thous<strong>and</strong> years ago (Botkin, 2001). The<br />

English garden, a much later development, represents the naturalistic (natural-like) idea of<br />

l<strong>and</strong>scape beauty, <strong>and</strong> in the early Renaissance the wilderness <strong>and</strong> the power of nature<br />

symbolized the power of God <strong>and</strong> sublime beauty.<br />

L<strong>and</strong>scape preferences studies usually adopt objective quantification or normative judgments<br />

methods (Burley, 2006, Panagopoulos, 2009). Unfortunately, such methodological approaches<br />

fail to account for people’s subjectivity, a core element that informs their spatial preferences <strong>and</strong><br />

evaluations of l<strong>and</strong>scapes, <strong>and</strong> have resulted in simplistic interpretations of human-place<br />

interactions (Ohta, 2001).<br />

According to Burke (1958) aesthetic of l<strong>and</strong>scapes can be distinguished as “Picturesque”<br />

(dominated from asymmetry, scenes, nostalgic); “Beautiful” (feeling that induce in us a sense of<br />

affection <strong>and</strong> tenderness or “Sublime” (that is a pleasure that arises, from pain or fear). A<br />

violent emotion can be caused from sublime l<strong>and</strong>scapes – especially with heightened spiritual<br />

feelings- <strong>and</strong> the elements of pathos <strong>and</strong> empathy exist in sublimity. During the second half of<br />

the 18th century the meaning of the word “sublime” shifted from rhetorical aesthetic to a<br />

psychological sense. The theory of “Peri Hipsous” had established the distinction between the<br />

elevated style beauty <strong>and</strong> the capacity to raise passion (Kaplan, 1987; Ramos & Panagopoulos,<br />

2007).<br />

Deserts can be valued aesthetically, if they are beautiful, sublime or picturesque. To identify<br />

whether a desert l<strong>and</strong>scape is beautiful or not, a set of variables is to be investigated. These<br />

variables might be physical or not, they can be recognized visually <strong>and</strong> non-visually using our<br />

senses. The aesthetic value of the desert l<strong>and</strong>scape can be recognized when the specialists <strong>and</strong><br />

the community recognize together, the attributes of the aesthetic variables (Lothian, 1999).<br />

Objective of the present study is to find how a person interprets desert l<strong>and</strong>scapes using<br />

phenomenological methodology.<br />

2. Methodology<br />

A questioner was used as research tool to measure l<strong>and</strong>scape preference. The participants were<br />

11 men <strong>and</strong> five women, three of whom participated in preliminary (test) interviews.<br />

Participants ranged between the ages of 19 <strong>and</strong> 65, possessed a high school <strong>and</strong> above education,<br />

originated from Portugal <strong>and</strong> related to l<strong>and</strong>scape planning <strong>and</strong> design. Thirty four color<br />

photographs of natural <strong>and</strong> designed desert l<strong>and</strong>scapes were selected from l<strong>and</strong>scape<br />

photographs. The interview style was semi-structured, beginning with basic open-ended<br />

questions were then made more specific.<br />

The questions were constructed following the principles described by Patton (1980): 1)<br />

Experience/Behavior (what a person does or has done); 2) Opinion/Value questions (aimed at<br />

underst<strong>and</strong>ing the subject's cognitive <strong>and</strong> interpretive processes); 3) Feelings (emotional<br />

responses to experiences <strong>and</strong> thoughts); 4) Aesthetic knowledge (factual information); <strong>and</strong> 5)<br />

Sensory Experience (what is seen, heard, touched, tasted, <strong>and</strong>/or smelled).<br />

The time frame of the questions may vary as necessary at the discretion of the interviewer.<br />

Typical questions for this study include the following: `What do you see on the photo' What<br />

would you feel if you were there' `Have you ever been to a place like this', What element of<br />

the l<strong>and</strong>scape attracted you Other sections of the questionnaire included psychophysical<br />

questions <strong>and</strong> other with sustainability perceptions <strong>and</strong> economic preferences.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Beldade & T. Panagopoulos 2010. Integrating esthetical <strong>and</strong> ecological values at the central Asia l<strong>and</strong>scape change<br />

342<br />

3. Results <strong>and</strong> Conclusions<br />

Regardless of their fragile ecosystems <strong>and</strong> stigma of l<strong>and</strong>scapes to avoid, decertified areas enjoy<br />

numerous human-based <strong>and</strong> natural attractions that are in some cases unique in natural world.<br />

Finding of this study well reveal that arid areas have a great number of potential capacities.<br />

Natural features like dunes, salt domes, erosion sculpts badl<strong>and</strong>s, salt lakes, fresh <strong>and</strong> saline<br />

water springs, <strong>and</strong> animals well- adapted to hard-to-live circumstances, are just few instances of<br />

wonderful natural phenomena one can see in an arid area. cactus <strong>and</strong> halophyte plants, from<br />

small bushes to 6-meter high shrubs, are other natural picturesque sceneries that only deserts<br />

can offer. Also, human-based attractions like historical monuments, type of architecture <strong>and</strong><br />

materials used for construction purposes, the ways desert-dwellers produce <strong>and</strong> subsist or deal<br />

with drought are other striking views that can attract tourists.<br />

Tabular data relating to the distribution of scenic ratings, <strong>and</strong> the indications of preference<br />

among observers, were generated using SPSS statistical software. After qualitative analysis, 7<br />

main categories of cognitive aspects of natural l<strong>and</strong>scapes were identified. Each l<strong>and</strong>scape<br />

evoked each participant's memories <strong>and</strong> background, <strong>and</strong> altered through these influences,<br />

imagination/association, impression, aesthetic judgments, <strong>and</strong> meaning <strong>and</strong> attractiveness of<br />

nature.<br />

All these played their roles in the evaluation of desert l<strong>and</strong>scapes (Figure 1). Each question<br />

caused that the participants remembered individual memories from their life <strong>and</strong> memories<br />

acquired through media or education. Participant’s background, hobbies <strong>and</strong> personal preference<br />

influenced the l<strong>and</strong>scape cognition. Time of day <strong>and</strong> season were also associated with questions<br />

related to first impressions about the l<strong>and</strong>scape. Imaginative activities <strong>and</strong> modifications<br />

concerning the l<strong>and</strong>scapes as well as the feeling of greatness of nature influence the answers.<br />

Also, because the participants were related to planning <strong>and</strong> design of l<strong>and</strong>scapes they judged the<br />

l<strong>and</strong>scape aesthetically depending on particular elements of the scene <strong>and</strong> they were able to<br />

underst<strong>and</strong> it as a photograph or as an actual scene.<br />

14<br />

Desert Topical Isl<strong>and</strong> Mountain Rural<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Liberty<br />

Discomfort<br />

Solitude<br />

Admiration<br />

Fear<br />

Enthusiasm<br />

Surprise<br />

Adventurous<br />

Curious<br />

Boring<br />

Sadness<br />

Harmony<br />

Asymmetry<br />

Simplicity<br />

Equilibrium<br />

Power<br />

Vastness<br />

Spiritual<br />

Obscure<br />

Incomprehensible<br />

Happiness<br />

Figure 1. Feelings evoked by each participant when evaluating different types of l<strong>and</strong>scapes.<br />

L<strong>and</strong>scape cognition with photographs showed dynamic <strong>and</strong> complicated interactions of various<br />

mental aspects during the course of each session. Participants had sufficient time for various<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Beldade & T. Panagopoulos 2010. Integrating esthetical <strong>and</strong> ecological values at the central Asia l<strong>and</strong>scape change<br />

343<br />

inner responses to arise naturally during his/her interaction with the l<strong>and</strong>scapes. However, most<br />

conventional quantitative studies of l<strong>and</strong>scape evaluation allow viewing times of only<br />

approximately few seconds per l<strong>and</strong>scape <strong>and</strong> with such a procedure researchers can deal only<br />

with participants' first impressions or fragmented memories (Ohta, 2001).<br />

Desert l<strong>and</strong>scapes show similar results of aesthetic value as a tropical isl<strong>and</strong> l<strong>and</strong>scape, while<br />

they provoke more fear, sense of power, vastness, sense of more spiritual, obscure, <strong>and</strong><br />

incomprehensible l<strong>and</strong>scape than any other compared. In all characteristics of a sublime<br />

l<strong>and</strong>scape, deserts <strong>and</strong> desert storms had the highest score. Also they had high scores in some<br />

attributes of the aesthetic variables for beautiful <strong>and</strong> picturesque l<strong>and</strong>scape like: harmony,<br />

balance <strong>and</strong> simplicity for beautiful, or asymmetry <strong>and</strong> pictorial value.<br />

A large majority of the individuals responded that whishes to buy a holiday house in desert<br />

l<strong>and</strong>scape offering till 200.000 euros, but 25% responded that is not wishing to have house in<br />

this kind of l<strong>and</strong>scape. 19% could spend more than 500 euros to travel in desert l<strong>and</strong>scape <strong>and</strong><br />

another 500 euros for subsistence.<br />

From the field survey it was found that majority of participants (56%) consider that the cost of<br />

reclamation practice at decertified areas should be paid from the residents of those areas or<br />

neighborhood areas because they provoked the desertification. Although, a 38% consider that<br />

desertification is a universal problem <strong>and</strong> is responsibility of everybody to avoid l<strong>and</strong>scape<br />

degradation.<br />

Only 4 out of 16 participants consider that the Mediterranean garden could be the desirable<br />

solution for tourism development in decertified l<strong>and</strong>scapes. Six prefer exotic species gardens<br />

(like oasis) 3 prefer Zen stile spaces (no vegetation) <strong>and</strong> two would rather prefer large loan areas<br />

with few trees. The large majority prefers to develop in those areas naturalistic l<strong>and</strong>scapes or<br />

with minimum human intervention, while none would like to see dominant tourism<br />

development projects to overtake in those l<strong>and</strong>scapes. A large number of participants (68%)<br />

requests that those areas to be preserved as of unique value as l<strong>and</strong>scapes <strong>and</strong> ecosystems.<br />

In Figure 2 can be seen the most preferable l<strong>and</strong>scape aesthetic elements that the participants<br />

consider as most attractive for decertified l<strong>and</strong>scapes. Human presence, luck of vegetation,<br />

geometric lines could be the less attractive elements in desert l<strong>and</strong>scapes, while water presence,<br />

color contrast <strong>and</strong> texture variability, wildlife, panoramic views <strong>and</strong> organic natural lines were<br />

the most attractive. Cultural elements, activities, dense vegetation like oasis, natural sounds <strong>and</strong><br />

silence could also be important to consider for tourism development sites <strong>and</strong> scenes.<br />

The current study contributes to this endeavor by exploring tourists’ perception of the desert<br />

l<strong>and</strong>scapes of central Asia (Aral Sea <strong>and</strong> Silk Road area); a l<strong>and</strong>scape that is suffering severe<br />

Aeolian erosion from dust storms <strong>and</strong> is undergoing various reclamation measures. Exploration<br />

of tourist’s l<strong>and</strong>scape meanings is informed by literature on the interpretation of space <strong>and</strong> place.<br />

Barnes & Duncan, (1992) makes a distinction between three dialectical structures of space,<br />

namely, spatial practices, representations of space <strong>and</strong> spaces of representation. Spatial practices<br />

manifest into social l<strong>and</strong>scapes over time. Representations of space are practices which organize<br />

<strong>and</strong> represent space, particularly through planning <strong>and</strong> design. Spaces of representation are<br />

spatialities ‘‘space as directly lived through its associated images <strong>and</strong> symbols” as understood<br />

by locals, tourists <strong>and</strong> tourism officials who compete for meanings, <strong>and</strong> uses.<br />

The main objective of this study was to create a better underst<strong>and</strong>ing of how a person interprets<br />

desert l<strong>and</strong>scapes, to examine public awareness <strong>and</strong> performance in the promotion of<br />

reclamation projects on decertified environment a to describe how to integrate l<strong>and</strong>scape<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Beldade & T. Panagopoulos 2010. Integrating esthetical <strong>and</strong> ecological values at the central Asia l<strong>and</strong>scape change<br />

344<br />

aesthetics into l<strong>and</strong>scape planning of the Central Asia Silk Road area. From the preliminary<br />

results of the first group of participants (planners <strong>and</strong> designers) it was found that the public<br />

have limited awareness <strong>and</strong> a poor underst<strong>and</strong>ing about sustainable development in global scale.<br />

Even that it was considered that desertification is a universal problem <strong>and</strong> is responsibility of<br />

everybody to avoid l<strong>and</strong>scape degradation, preserve <strong>and</strong> promote the desert l<strong>and</strong>scapes <strong>and</strong><br />

ecosystems as of unique value. Also, it was found that ecological aesthetics pleasure could be<br />

derived from knowing on how part of the l<strong>and</strong>scape relate to the whole.<br />

100%<br />

90%<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

Human presence<br />

Water presence<br />

No attraction Indifferent Attracts High attraction<br />

Colour contrast<br />

Texture variability<br />

Natural lines<br />

Geometric lines<br />

Cultural elements<br />

Wildlife<br />

Movement<br />

Panoramic view<br />

Light<br />

Shadows<br />

Contained space<br />

large space<br />

Natural sounds<br />

Silence<br />

Disperse vegetation<br />

Dense vegetation<br />

Accessibilities<br />

Planes<br />

High slopes<br />

Heat<br />

Cold<br />

Figure 2. L<strong>and</strong>scape aesthetic elements that may attract to a decertified l<strong>and</strong>scape.<br />

Opinion of the authors is that in the wake of global climate change, society has to address the<br />

dynamic interactions between an increasingly changing environment <strong>and</strong> its politically <strong>and</strong><br />

culturally contested spatial development in decertified l<strong>and</strong>scapes.<br />

Acknowledgements<br />

This work was financed from the European Union for the “Long Term Ecological Research<br />

Program for Monitoring Aeolian Soil Erosion in Central Asia” (CALTER), FP6-2003-INCO-<br />

Russia+NIS (project Nº 516721). The authors also want to gratefully acknowledge the Centro<br />

de Investigação sobre Espaço e Organizações (CIEO) for having provided the conditions to<br />

publish this work.<br />

References<br />

Botkin, D.B., 2001. An ecologist’s ideas about l<strong>and</strong>scape beauty: Beauty in art <strong>and</strong> scenery as<br />

influenced by science <strong>and</strong> ideology. In Sheppard, S.R.J., & Harshaw H.W. (eds.). <strong>Forest</strong>s<br />

<strong>and</strong> l<strong>and</strong>scapes: Linking ecology, sustainability <strong>and</strong> aesthetics, CABI Publishing,<br />

Wallingford, United Kingdom, pp. 111-123.<br />

Brunson, M., & Shelby, B., 1992. Assessing recreational <strong>and</strong> scenic quality: How does new<br />

forestry rate Journal of <strong>Forest</strong>ry, 90(7), 37–41.<br />

Burke, E., 1958. A phylosophical inquiry into the origins of the sublime <strong>and</strong> beautiful.<br />

Routledge, London.<br />

Burley, B.J., 2006. A quantitative method to assess aesthetic/environmental quality for spatial<br />

surface mine planning <strong>and</strong> design applications. WSEAS Transactions on Environment <strong>and</strong><br />

Development, 2, 524-529.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Beldade & T. Panagopoulos 2010. Integrating esthetical <strong>and</strong> ecological values at the central Asia l<strong>and</strong>scape change<br />

345<br />

Cambers, G., 2009. Caribbean beach changes <strong>and</strong> climate change adaptation. Aquatic Ecosystem<br />

Health <strong>and</strong> Management, 12(2), 168–176.<br />

Canter, L.W., 1996. Environmental Impact Assessment. McGRaw-Hill International Editions,<br />

Singapore.<br />

Carlson, A., 2002. Aesthetics <strong>and</strong> the Environment: The Appreciation of Nature, Art <strong>and</strong><br />

Architecture. Routledge, New York.<br />

Kaplan, S., 1987. Aesthetics, affect, <strong>and</strong> cognition. Environmental preference from an<br />

evolutionary perspective. Environment <strong>and</strong> Behaviour, 19:3-32.<br />

Kemal, S., Gaskell, I., 1993. L<strong>and</strong>scape, Natural Beauty <strong>and</strong> the Arts. Cambridge University<br />

Press. Cambridge.<br />

Lothian, A.., 1999. L<strong>and</strong>scape <strong>and</strong> the philosophy of aesthetics: is l<strong>and</strong>scape quality inherent in<br />

the l<strong>and</strong>scape or in the eye of the beholder L<strong>and</strong>scape Urban Planning, 44, 177-198.<br />

Ohta, H., 2001. A phenomenological approach to natural l<strong>and</strong>scape cognition. Journal of<br />

Environmental Psychology, 21: 387-403.<br />

Panagopoulos T., 2009. Linking forestry, sustainability <strong>and</strong> aesthetics. Ecological Economics,<br />

Vol. 68: 2485-2489.<br />

Patton, M. Q., 1980. Qualitative Evaluation Methods. SAGE Publications, Beverly Hills, U.S.A.<br />

Ramos, B. & Panagopoulos, T., 2007. Integrating aesthetic <strong>and</strong> sustainable principles in stream<br />

reclamation projects. WSEAS Transactions on Environment <strong>and</strong> Development, 3, 189-195.<br />

Thorne, J.F. & Huang, C.S., 1991. Toward a l<strong>and</strong>scape ecological aesthetic: methodologies for<br />

designers <strong>and</strong> planners. L<strong>and</strong>scape <strong>and</strong> Urban Planning 21, 61-79.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Carvalho & A.R. de Freitas 2010. The impact of legislation on the dynamics of l<strong>and</strong> use the River Basin Cará-Cará<br />

346<br />

The impact of legislation on the dynamics of l<strong>and</strong> use the River Basin<br />

Cará-Cará, Ponta Grossa-PR/Brazil, in period from 1980 to 2007<br />

Silvia Méri Carvalho 1* & Andreza Rocha de Freitas 2<br />

1 Universidade Estadual de Ponta Grossa – UEPG, Ponta Grossa, Paraná, Brazil<br />

2 Universidade Estadual do Centro-Oeste – UNICENTRO, Irati, Paraná, Brazil<br />

Abstract<br />

The human ownership of area causes transformations that can be accompanied <strong>and</strong> identified by<br />

means of l<strong>and</strong> use, delineation of areas of environmental conflicts <strong>and</strong> categories of hemeroby.<br />

The objective of this work is examine the dynamics of occupation of the l<strong>and</strong> use in Cará-Cará<br />

hydrographic basin located in municipal district of Ponta Grossa, in Parana State – Brazil,<br />

between the years 1980 <strong>and</strong> 2007, <strong>and</strong> the relation of laws in force with the l<strong>and</strong> use. It was<br />

necessary to search <strong>and</strong> draw up maps of relevant legislation, of slope <strong>and</strong> l<strong>and</strong> use that were<br />

overlay until they reach to maps synthesis of conflicts l<strong>and</strong> use. As for l<strong>and</strong> use, the urban class<br />

increased (121.83%) due to the increase in population. Areas that are in conflict regarding the<br />

use represent 21.05%. Classes of hemeroby showed that the l<strong>and</strong>scapes more artificial than<br />

natural occupy 41.94%. It was concluded that the changes in Cará-Cará hydrographic basin<br />

were motivated by guidelines of the Municipal Master Plans.<br />

Keywords: Hydrographic basin, Environmental conflicts, L<strong>and</strong> use, Hemeroby<br />

1. Introduction<br />

The changes caused by human activities in space can be monitored <strong>and</strong> verified through a<br />

survey of l<strong>and</strong> use, the delimitation of areas of environmental conflicts <strong>and</strong> the identification of<br />

areas that suffered most changes in their natural characteristics.<br />

When dealing with topics <strong>and</strong> themes used in environmental planning, Santos (2004, p. 97)<br />

states that l<strong>and</strong> use is a basic theme, it "portrays the human activities that can create pressure on<br />

the natural elements" which describes "not only the current situation, but recent changes <strong>and</strong><br />

history of occupation of the study area".<br />

According to Clawson & Stewart (1965), l<strong>and</strong> use refers to human activity on l<strong>and</strong>, which is<br />

directly connected to earth. With the same thought Campbell (1997) states that l<strong>and</strong> use are the<br />

human activities on l<strong>and</strong> held as needed, <strong>and</strong> the results of these activities are physical changes<br />

that transform the environment.<br />

There is need for strategies to maintain balance <strong>and</strong> dynamics of natural ecosystems present, for<br />

example in hydrographic basin. These strategies can be implemented <strong>and</strong> controlled by<br />

legislation, such as the Municipal Master Plan <strong>and</strong> the <strong>Forest</strong> Code to address the Permanent<br />

Preservation Areas. The occupation of these areas with other uses generates so-called<br />

environmental conflicts.<br />

According to Rocha (1997) environmental conflicts in l<strong>and</strong> use occur when agricultural<br />

* Corresponding author.<br />

Email address: silviameri@brturbo.com.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Carvalho & A.R. de Freitas 2010. The impact of legislation on the dynamics of l<strong>and</strong> use the River Basin Cará-Cará<br />

347<br />

activities are practiced in inappropriate areas, <strong>and</strong> these activities are most responsible for the<br />

erosion, siltation of rivers, floods <strong>and</strong> droughts.<br />

The effects of the use <strong>and</strong> occupation of areas protected by law may cause the deterioration of<br />

the environment, emerging, so-called environmental conflicts in l<strong>and</strong> use (FERNANDES NETO<br />

& ROBAIANA, 2005).<br />

To study the effects caused by human action on the various biological systems, according to<br />

Dueñas (2004), it is necessary to develop a systematic method, comparative <strong>and</strong> qualitative,<br />

which permits the effect of human disturbance on the different elements of ecosystems.<br />

Arise, thus concepts that serve as the basis for monitoring the developments <strong>and</strong> changes in l<strong>and</strong><br />

use caused. The concept of hemeroby is one. This term was suggested by Jalas (1953) which<br />

determines the degree of alteration of l<strong>and</strong>scapes, in other words, the degree of naturalness <strong>and</strong><br />

artificiality of the medium, considering the following classification: Ahemerobe - natural<br />

l<strong>and</strong>scapes or small human interference, such as rainforest <strong>and</strong> gallery forest; Oligohemerobe -<br />

l<strong>and</strong>scapes more natural than artificial, like dirty fields used for livestock; Mesohemerobe -<br />

l<strong>and</strong>scapes more artificial than natural, such as reforestation, <strong>and</strong> Euhemerobe - artificial<br />

l<strong>and</strong>scapes, such as cultivated areas <strong>and</strong> urban area.<br />

The anthropogenic impact can be evaluated through studies that show where are the areas most<br />

degraded <strong>and</strong> modified, primarily through the analysis <strong>and</strong> temporal-spatial representation of<br />

l<strong>and</strong> use.<br />

Studies of this nature are part of environmental planning because they provide information<br />

needed to develop strategies <strong>and</strong> actions to mitigate the impacts caused by human interference.<br />

It is a study that could help in the delimitation of areas to be occupied or eventually recovered.<br />

Therefore, the general aim of this study is to analyze the impact of the legislation in force in the<br />

dynamics of occupation of l<strong>and</strong> use in the period 1980 to 2007. Thus, it was necessary to draw<br />

up maps of l<strong>and</strong> use, raise the relevant legislation for the period studied <strong>and</strong> elaborate maps of<br />

environmental conflicts in l<strong>and</strong> use.<br />

2. Methodology<br />

The input, storage, processing <strong>and</strong> output data was performed using the software SPRING,<br />

version 4.3.3, developed by the National Institute for Space Research/Division of Image<br />

Processing (INPE / DPI, 1999).<br />

Based on the topographical maps, in digital media, prepared by the Directorate of Geographic<br />

Services (DSG) Army (1980), scale 1:50,000, sheet SG.22-XC-II/2 (Ponta Grossa), data were<br />

taken drainage, contour, perimeter of the basin, roads <strong>and</strong> highways.<br />

After scanning the drainage network buffers were created to set the Permanent Preservation<br />

Areas (PPAs) along the rivers <strong>and</strong> springs, as provided by the <strong>Forest</strong>ry Code (BRASIL, 1965).<br />

For the preparation of maps of l<strong>and</strong> use was used aerial photography from 1980 (Institute of<br />

L<strong>and</strong> <strong>and</strong> Cartography of Parana - ITC) in the scale of 1:25,000, <strong>and</strong> the image CBERS2, March<br />

2007. The composition of b<strong>and</strong>s adopted in the present work was the R (3) G (4) B (2). The<br />

method adopted for the classification of the images was the Supervised Classification by<br />

algorithm MaxVer.<br />

Subclasses themes of l<strong>and</strong> use were adapted from the Technical Manual of L<strong>and</strong> Use from the<br />

Brazilian Institute of Geography <strong>and</strong> Statistics - IBGE (IBGE, 2006). Subclasses were defined<br />

as follows: urbanized area, cultivation (including temporary <strong>and</strong> permanent crops), forest,<br />

grassl<strong>and</strong> <strong>and</strong> water body. In addition to the subclasses suggested was adopted by the IBGE<br />

class industrial area <strong>and</strong> reforestation.<br />

After completion of all maps was used superposition method (Santos, 2004) with the binary<br />

crossing until reach the intermediate maps that were in turn overlapped.<br />

To identify the degree of naturalness/artificiality in the Cará-Cará basin used the scenarios<br />

focusing on l<strong>and</strong> use from 1980 to 2007.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Carvalho & A.R. de Freitas 2010. The impact of legislation on the dynamics of l<strong>and</strong> use the River Basin Cará-Cará<br />

348<br />

After preparation of scenarios of l<strong>and</strong> use adopted the concept <strong>and</strong> classification of hemerobia<br />

of Jalas (1953) in the preparation of letters of artificiality in the Cará-Cará basin. The Categories<br />

adopted are: ahemerobe, oligohemerobe, mesohemerobe <strong>and</strong> euhemerobe.<br />

3. Result<br />

Two scenarios l<strong>and</strong> use were constructed for Cará-Cará basin for the years 1980 <strong>and</strong> 2007<br />

(Figure 1), where they were identified in the first scenario five classes of l<strong>and</strong> use <strong>and</strong> seven<br />

classes in the second (Table 1).<br />

Figure 1. Maps of l<strong>and</strong> use of Cará-Cará Hydrographic Basin of 1980 <strong>and</strong> 2007<br />

Table 1. Quantification of l<strong>and</strong> use classes<br />

Classes<br />

1980 2007<br />

Area (ha) % Area (ha) %<br />

Industrial area - - 314.59 4.30<br />

Urbanized area 343.72 4.70 762.48 10.42<br />

Campestral 3,802.30 51.96 3,069.06 41.94<br />

Industrial area<br />

Continental water - - 2.64 0.04<br />

body<br />

Cultivation 2,126.42 29.06 2,423.79 33.12<br />

<strong>Forest</strong>ry 363.42 4.97 274.36 3.75<br />

Reforestation 681.66 9.31 470.60 6.43<br />

Total 7,317.52 100 7,317.52 100<br />

To map the areas of environmental conflicts, we used the proposal to Beltrame (1994)<br />

represented by the classes corresponding use <strong>and</strong> areas over-used <strong>and</strong> under-utilized. In this<br />

study, the corresponding use areas are being used as the relevant legislation. The over-used<br />

areas are those where the activities are being carried out in permanent preservation areas or not<br />

provided for in the Zoning Plan. Underutilized areas in the Cará-Cará Hydrographic basin are<br />

represented by areas for urbanization <strong>and</strong> industry, not yet occupied by these activities (Figure<br />

2) (Table 2).<br />

Table 2. Environmental conflicts in the Cará-Cará hydrographic basin<br />

Classes<br />

1980 2007<br />

Area (ha) % Area (ha) %<br />

Over-used 409.62 34.26 153.39 9.96<br />

Underused 786.15 65.74 1,387.06 90.04<br />

Total 1,195.77 100 1,540.45 100<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Carvalho & A.R. de Freitas 2010. The impact of legislation on the dynamics of l<strong>and</strong> use the River Basin Cará-Cará<br />

349<br />

Figure 2. Maps of environmental conflicts of Cará-Cará Hydrographic Basin of 1980 <strong>and</strong> 2007.<br />

As a result of the maps of artificiality of the medium we obtained two letters of hemerobia for<br />

Cará-Cará Hydrographic basin dated of 1980 <strong>and</strong> 2007 (Figure 3), which were analyzed<br />

according to the degree of anthropogenic interference existing.<br />

Figure 3. Maps of hemerobia of Cará-Cará Hydrographic Basin of 1980 <strong>and</strong> 2007.<br />

Were identified <strong>and</strong> mapped four classes of hemerobia: ahemerobe, oligohemerobe,<br />

mesohemerobe <strong>and</strong> euhemerobe (Table 3).<br />

4. Discussion<br />

Table 3. Quantification of hemerobe classes of Cará-Cará Hydrographic Basin<br />

Classes<br />

1980 2007<br />

Area (ha) % Area (ha) %<br />

Ahemerobe 363.42 4.97 274.36 3.75<br />

Oligohemerobe 3,802.30 51.96 3,069.06 41.94<br />

Mesohemerobe 2,808.08 38.37 2,897.03 39.59<br />

Euhemerobe 343.72 4.70 1,077.07 14.72<br />

Total 7,315.52 100 7,315.52 100<br />

Between the years 1980 <strong>and</strong> 2007, three classes of l<strong>and</strong> use had exp<strong>and</strong>ed, namely: industrial<br />

area, urbanized area <strong>and</strong> cultivation. In 1980 the industrial district was already defined in the<br />

Cará-Cará basin, however there were no industries located. The area dedicated to industrial<br />

activities at this time was occupied by reforestation (Pinus spp.) raw material for wood<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Carvalho & A.R. de Freitas 2010. The impact of legislation on the dynamics of l<strong>and</strong> use the River Basin Cará-Cará<br />

350<br />

industries located in the vicinity.<br />

The urbanized area, was the class with the biggest increase during the study period (121.83%)<br />

<strong>and</strong> is located in the northwest portion of the Cará-Cará basin. These areas were employed due<br />

the implementation of subdivisions <strong>and</strong> areas intended for urban expansion planned in the city<br />

plan.<br />

The cultivation class, represents areas occupied by temporary <strong>and</strong> perennial crops, mainly<br />

soybean planting, followed by corn <strong>and</strong> wheat. These areas are located mainly in central <strong>and</strong><br />

have advanced to the northeast of the basin, using areas that were once occupied by the field <strong>and</strong><br />

forests. For the period studied there was an increase of 13.98% in total area.<br />

The areas occupied by native vegetation, represented by the forest <strong>and</strong> grassl<strong>and</strong> were the most<br />

had a reduction in area, <strong>and</strong> the forest showed a decrease of 24.51% for the period analyzed, due<br />

to the expansion of urban activities, industry <strong>and</strong> farming . The forest areas are distributed on<br />

the banks of some rivers, especially in the southern portion near the mouth of the Cará-Cará<br />

River <strong>and</strong> north in the area belonging to the 13th BIB, where there was the recomposition of<br />

forest <strong>and</strong> the area is used in training the army. The campestral class was affected by the<br />

occupation of other activities, decreasing by 19.28% the area occupied between 1980 <strong>and</strong> 2007.<br />

The reforestation class, representing the areas reforested with eucalyptus (Eucalyptus spp) <strong>and</strong><br />

pine (Pinus spp), was unstable between the years 1980 <strong>and</strong> 2007. This occurred because in the<br />

80, industrial output surpasses agriculture, <strong>and</strong> this transition the so-called traditional industries<br />

(textiles, wood, food, furniture, etc..) lose relative importance in the economy of the state of<br />

Parana. The wood <strong>and</strong> textile sectors were the most reduced their participation in the GDP of<br />

the state (Migliorini, 2006).<br />

As can be observed during the study period, the class with the biggest increase was the<br />

urbanized area (121.83%) due to population increases <strong>and</strong>, consequently, the expansion of the<br />

town of Ponta Grossa, in the east of the Cará-Cará Hydrographic Basin. The increased area<br />

occupied by urban activities took place, not only by increasing the population of the city, but<br />

also by urban voids used for speculation.<br />

In relation to environmental conflicts, over-used areas characterized by the illegal occupation in<br />

areas that should be protected or not built that were more varied. Between the years 1980 <strong>and</strong><br />

2001 the variation was 7.99% this is due to the establishment of areas of Funds Master Plan for<br />

the Vale of Ponta Grossa that were being occupied by urban <strong>and</strong> agricultural activities.<br />

Areas of environmental conflict deemed over-use continues to increase, so even if incipient, on<br />

permanent preservation areas <strong>and</strong> not building l<strong>and</strong> in the central portion of the basin.<br />

The other conflicting class is regarded as underutilized, where the areas that should be being<br />

used by certain activities have not yet been occupied. In the period studied the class that this<br />

was more varied <strong>and</strong> increased (76.44%) due to changes in local laws, especially in residential<br />

zoning <strong>and</strong> industrial <strong>and</strong> extinction of Special L<strong>and</strong>scape Areas.<br />

The maps of environmental conflicts assist in identifying areas where there is greater human<br />

interference in Cará-Cará Hydrographic Basin, making it more natural <strong>and</strong> less artificial. To<br />

facilitate the analysis of the anthropogenic impact were prepared letters of hemerobia.<br />

The ahemerobe class corresponds to remnants of Araucaria forests in different successional<br />

stages. Ahemerobe areas are occupied by urban activities, industry <strong>and</strong> culture, representing a<br />

decrease of 24.51% of the basin area, being more significant in the northern (13 º BIB) <strong>and</strong> on<br />

the banks of rivers in the south.<br />

The oligohemerobe class, represented by dirty fields used for grazing cattle, also showed<br />

decreased area occupied (19.28%) due to the advancement of industrial, urban <strong>and</strong> agricultural<br />

l<strong>and</strong>s.<br />

Unlike previous classes, the classes mesohemeorbe <strong>and</strong> euhemerobe had exp<strong>and</strong>ed between<br />

1980 <strong>and</strong> 2007. The mesohemerobe class is characterized largely by corn, wheat <strong>and</strong> soybeans<br />

at EMBRAPA <strong>and</strong> reforestation of pine (Pinus spp) <strong>and</strong> eucalyptus (Eucalyptus spp) in IAPAR.<br />

It may be noted that this class has exp<strong>and</strong>ed only 3.17%.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Carvalho & A.R. de Freitas 2010. The impact of legislation on the dynamics of l<strong>and</strong> use the River Basin Cará-Cará<br />

351<br />

The euhemerobe class represents the areas more artificial of Cará-Cará Hydrographic Basin, or<br />

those occupied by industrial <strong>and</strong> urban activities. The class has advanced over 200% compared<br />

to 1980.<br />

The changes in the use of l<strong>and</strong>, which formed the basis for determining the hemerobe classes<br />

were given due to changes in territorial planning of the city of Ponta Grossa.<br />

References<br />

Beltrame, Â. V. Diagnóstico do meio físico de Bacias Hidrográficas: modelo e aplicação.<br />

Florianópolis: Ed. da UFSC, 1994.<br />

Brasil, Lei n.º 4.771, de 15 de setembro de 1965. Institui o Código Florestal.<br />

Campbell, J. B. L<strong>and</strong> Use <strong>and</strong> Cover Inventory. In: Manual of Photographic Interpretation. 2a<br />

ed. USA: ASPRS, 1997. 335 a 360p.<br />

Clawson, M.; STEWART, S. I. L<strong>and</strong> use information - A critical survey of us statistics<br />

including possibilites for greater uniformity. Baltimore, Md: The John Hopkins Press for<br />

Resources for the Future, Inc, 1965.<br />

Dsg/Codepar. Folha de Ponta Grossa. Rio de Janeiro: [s. n.], 1980. 1 mapa: SG.22-X-C-II-2.<br />

Escala 1:50.000; Lat. 25º 00’ – 25º 15’; Long. 50º 00’ – 50º 15’. Região Sul do Brasil.<br />

Dueñas, W. A. M. Estudio integrado del grado de antropización (INRA) a escala del paisaje:<br />

propuesta metodológica y evaluación. IASCP, Colômbia, 2004.Disponível em:<br />

Consultado em 28/06/2007.<br />

Fern<strong>and</strong>es Neto, S.; Robaina, L. E. Conflito de Uso da Terra – Oeste do RS. In: SIMPÓSIO<br />

BRASILEIRO DE GEOGRAFIA FÍSICA APLICADA, 11, 2005, São Paulo. Anais…<br />

São Paulo: USP, 2005. p. 2728-2741.<br />

IBGE. Instituto Brasileiro de Geografia e Estatística. Manual Técnico de Uso da Terra. Série<br />

Manuais Técnicos em Geociências, Rio de Janeiro, n.º 7, 2 ed., 2006.<br />

INPE-DPI. SPRING, Manual do usuário, São José dos Campos, 1999.<br />

(http://www.inpe.br/spring).<br />

Jalas, J. Hemerokorit ja hemerobit.- Luonnon Tutkija, 1953, 57, p. 12-16.<br />

Migliorini, S. M. S. Indústria paranaense: formação, transformação econômica a partir da<br />

década de 1960 e distribuição espacial da indústria no início do século XXI. Revista<br />

Eletrônica Geografar, Curitiba, v.1, n.1, p. 62-80, jul./dez. 2006<br />

Rocha, J. S. M. Manual de Projetos Ambientais. Imprensa Universitária, Santa Maria, 1997.<br />

Santos, R. F. Planejamento Ambiental: teoria e prática. São Paulo: Oficina de Textos, 2004.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Quantitative assessment of temporal dynamics in altitudinal-driven ecotones<br />

352<br />

Quantitative assessment of temporal dynamics in altitudinal-driven<br />

ecotones in a section of Valtellina Italian Alps<br />

Ramón Alberto Diaz-Varela 1* , Silvia Calvo-Iglesias 2 , Michele Meroni 3 & Roberto<br />

Colombo 3<br />

1 University of Santiago de Compostela, Spain<br />

2 University of Vigo, Spain<br />

3 University of Milano-Biccoca, Italy<br />

Abstract<br />

Mountain ecotones are sensitive to climate <strong>and</strong> global change <strong>and</strong> their historical dynamics can<br />

be used as a record <strong>and</strong> indicator of such events. Nevertheless there are relatively few studies<br />

aiming at the quantification of their dynamics in a spatial explicit way at a detailed scale. In this<br />

work we quantified the altitudinal shifts <strong>and</strong> spatial pattern of mountain ecotones. We applied a<br />

novel procedure to delineate the current <strong>and</strong> former location of three characteristic mountain<br />

ecotones, formalised as forest, tree <strong>and</strong> tundra lines in a section of Valtellina (Italian Alps). We<br />

estimated the medians of overall decadal altitude increments in 25 m for forest line, 13 m for<br />

tree line <strong>and</strong> 11 m for tundra line. The method also allowed us to differentiate between ecotone<br />

advance <strong>and</strong> retreat events. We also conducted an analysis of vegetation patches morphology at<br />

the ecotone locations, which showed significant implications in their dynamics.<br />

Keywords: Mountain ecotones, Alps, <strong>Global</strong> change, Spatially explicit model, Digital elevation<br />

model<br />

1. Introduction<br />

Mountain <strong>and</strong> boreal ecosystems are among the environments most susceptible to climate <strong>and</strong><br />

l<strong>and</strong> use changes (Theurillat <strong>and</strong> Guisan 2001; Didier 2001). Both phenomena have the general<br />

effect of raising of altitude driven ecotones (Didier 2001; Körner 1998; 1999). Despite the<br />

considerable number of works dealing with the definition <strong>and</strong> environmental implications of the<br />

upper limits of mountain vegetation, there are few works that focus on its spatially explicit<br />

delineation. Besides, the literature offers many examples of woodl<strong>and</strong> upper limits while less<br />

attention has been paid to alpine treeless vegetation <strong>and</strong> its limits with non-vegetated rocky <strong>and</strong><br />

nival habitats.<br />

The main aim of this paper is to analyse the spatial <strong>and</strong> temporal dynamics of forest line, tree<br />

line <strong>and</strong> tundra line ecotones in alpine areas <strong>and</strong> to evaluate their fluctuations in the context of<br />

climatic <strong>and</strong> l<strong>and</strong> use change.<br />

We tested the method in the Val Masino catchment in the Central Alps (province of Sondrio,<br />

Lombardy region, Italy). The study area comprises two sub-catchments (Valle dei Bagni <strong>and</strong><br />

Val di Mello) of the Masino creek basin, comprising 83.4 km 2 of surface <strong>and</strong> with a range of<br />

altitudes from 909 to 3432 m a.s.l. (Fig. 1). Vegetation types <strong>and</strong> patterns are closely related to<br />

altitudinal gradients. The uppermost sectors are occupied by glaciers, bare or sparse vegetated<br />

surfaces on rocky habitats such as screes, rocky slopes or cliffs. Below this belt different kinds<br />

* Corresponding author. Tel.: +34 982 285 900 ext. 22482 - Fax: +34 982 285 985<br />

Email address: ramon.diaz@usc.es<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Quantitative assessment of temporal dynamics in altitudinal-driven ecotones<br />

353<br />

of herbaceous formations, dwarf scrubs <strong>and</strong> thickets take place. Woodl<strong>and</strong>s with variable<br />

density of scattered <strong>and</strong> often stunted individuals of pine (Pinus sylvestris, P. mugo), spruce<br />

(Picea abies) <strong>and</strong> larch (Larix decidua) occur between this treeless area <strong>and</strong> dense forest<br />

formations. These species also dominate the high altitude forests, while the lower parts are<br />

covered by different tree formations dominated by deciduous broadleaved species. Hay<br />

meadows are the most common agricultural l<strong>and</strong>, occupying only a small surface at valley<br />

bottoms, while arable l<strong>and</strong> is virtually absent in the area.<br />

2. Methodology<br />

We started with the stereoscopic analysis <strong>and</strong> visual interpretation of aerial photos from 1954<br />

<strong>and</strong> 2003 to generate a raster map with the distribution of close forest, scattered trees <strong>and</strong> alpine<br />

tundra for each of the dates. We then used a spatial overlay technique to analyse the temporal<br />

change of the three classes. The overlay outputs were arranged in a 44 contingency matrix<br />

using as input the 1954 (columns) <strong>and</strong> 2003 (rows) raster maps, <strong>and</strong> considering 4 classes,<br />

namely: “Dense forest”, “Scattered trees”, “Tundra” <strong>and</strong> “Other”. Overall Kappa <strong>and</strong> Kappa<br />

indices for each l<strong>and</strong> cover class (Cohen 1960) were used as estimators of the rate of change, as<br />

suggested in other studies (Calvo-Iglesias et al. 2006).<br />

We then applied an algorithm developed by Diaz Varela et al. (2010) to identify the uppermost<br />

pixel of targeted ecotones for each of these maps. The method used to delineate forest line, tree<br />

line <strong>and</strong> tundra line ecotones follows the general definition of tree line by Körner (1999) <strong>and</strong><br />

aims at performing the automatic discrimination of elements (outpost) with an uppermost<br />

extreme position on a certain slope. Outposts corresponded to the uppermost pixels of each of<br />

the three l<strong>and</strong> cover classes <strong>and</strong> they were used to represent <strong>and</strong> map the ecotones. That is, for<br />

the l<strong>and</strong> cover class “forest”, outposts define the forest line; the “dense forest” plus “scattered<br />

trees” class outposts define the tree line <strong>and</strong> the “tundra” class outposts define the tundra line.<br />

An outpost is herein defined according to this example regarding trees: an operator in the field<br />

will define a tree as belonging to the tree line if he encounters no other tree while walking<br />

upward in the direction of maximum slope for as long as possible (i.e. until a summit is reached).<br />

Therefore, the trajectory from a given point will move upward following a maximum slope path<br />

until it encounters another element of the same class or finishes at a summit large enough to<br />

define a slope or series of slopes with similar exposure at a certain scale (cf. fig 2). Extending<br />

this concept from single trees to a the maps of forests, trees <strong>and</strong> tundra, the algorithm was<br />

implemented in ITT-IDLTM V. 6.3 to label the cells of a binary l<strong>and</strong> cover map (in this case<br />

tree – no tree) as belonging to the ecotone or not (see Diaz-Varela et al. 2010 for a more<br />

thorough description of the method).<br />

The temporal evolution of the different ecotones was analysed by extending the algorithm to a<br />

diachronic map. For the identification of ecotone altitude advances we ran upward flow paths<br />

from any upslope outpost pixel for 1954 until a 2003 upslope outpost pixel or a summit was<br />

reached. Altitude retreats were computed with the same scheme but using 2003 <strong>and</strong> 1954<br />

ecotone pixels as starting <strong>and</strong> end points, respectively. Therefore, where the ecotone location in<br />

1954 appeared in a higher position than in 2003 along a given maximum slope line, it was<br />

possible to trace <strong>and</strong> quantify the retreat or negative altitudinal shift.<br />

To characterise morphologically the spatial pattern of the outposts for the three ecotones in the<br />

period of reference <strong>and</strong> to assess changes in vegetation dispersion strategies, we used the<br />

GUIDOS software (Graphical User Interface for the Description of image Objects <strong>and</strong> their<br />

Shapes). This software was recently developed for morphological spatial pattern analysis<br />

(MSPA) of forest functional connectivity in the context of biodiversity studies (Vogt et al. 2008,<br />

Vogt et al., 2009). The output of the MSPA is a raster layer where patch cells are assigned to<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Quantitative assessment of temporal dynamics in altitudinal-driven ecotones<br />

354<br />

seven morphological spatial pattern classes: edge, core, perforated, islets, bridge, loop <strong>and</strong><br />

branch. We ran the software using the default parameters for the computation of MSPA,<br />

considering an edge width of 20 m.<br />

3. Results<br />

For the period analysed, tundra remained mostly stable while forest <strong>and</strong> scattered tree classes<br />

experienced more intense dynamics, as indicated by kappa values of Table 1. From the<br />

contingency matrix shown in this table, we observed that tundra also experienced some<br />

expansion colonising rocky areas <strong>and</strong> bare soils while in other areas it was colonised by<br />

scattered trees. The increase of dense forest occurred at the expenses of areas occupied by<br />

scattered trees in 1954.<br />

The totals <strong>and</strong> altitudinal distribution of the outpost is shown in Fig. 3. The forest line was<br />

located between 1350 <strong>and</strong> 1950 m, rarely reaching more than 2000 m. The tree line showed a<br />

slightly elevated position (around 100-150 m higher) on the slopes, while the tundra line lay<br />

clearly much higher, with maxima up to 2600 m.<br />

Figure 4 shows the altitude distribution, along with the horizontal <strong>and</strong> vertical increments of<br />

outpost altitudinal shifts, distinguishing positive (advance) <strong>and</strong> negative (retreats) altitude<br />

increments. A total of 225, 352 <strong>and</strong> 645 paths showing positive altitudinal shifts were recorded<br />

for forest line, tree line <strong>and</strong> tundra line respectively, with median values of 26, 17 <strong>and</strong> 15 m of<br />

decadal altitude increment respectively. Retreats were far less common than advances for the<br />

three ecotones under analysis. Putting together advances <strong>and</strong> retreats gives an overall view of<br />

the altitudinal shifts: the forest, tree <strong>and</strong> tundra lines showed a net advance, with medians of 25,<br />

13 <strong>and</strong> 11 m of altitude shifts respectively. Altitude plays a major role in the occurrence <strong>and</strong><br />

magnitude of the shifts. Advances of forest line took place mainly from 1300 to 1700 m. The<br />

altitude advance magnitude shows a clear trend to decrease with the rise in altitude. Thus, shifts<br />

that started from lower positions attained the most important advances, up to 600 m, while those<br />

starting near the top of the altitudinal limit for the class reached much lower values (e.g. less<br />

than 20 m on average in the interval 1900-2000). This tendency is less clear in the case of tree<br />

line, which showed more regular behaviour along its altitudinal range <strong>and</strong> altitudinal shifts<br />

which were generally lower than 200-300 m. The tundra line advance also shows the same trend<br />

to decrease its magnitude proportional to the altitude gradient, starting from large shifts (400-<br />

800 m) occurring at the lower locations (1600-1900 m) to small ones (less than 50 m) near its<br />

altitudinal limit (2700-2800 m). Retreat paths showed less clear patterns in terms of variation of<br />

magnitude along the altitudinal gradient.<br />

Similarly to l<strong>and</strong> cover dynamics, the spatial pattern of tundra experienced little variation (cf.<br />

Fig. 5, being characterised by a high amount of core <strong>and</strong> little proportion of edge <strong>and</strong> other<br />

morphological classes, though there is some increase of branches <strong>and</strong> islets reflecting the<br />

expansion <strong>and</strong> colonisation of new areas. Scattered trees <strong>and</strong> forest classes experienced an<br />

increase in core <strong>and</strong> decrease in edge <strong>and</strong> branches, as a result of their expansion they became<br />

more compact.<br />

4. Discussion<br />

The application of the method on multi-date l<strong>and</strong> cover datasets (1954 <strong>and</strong> 2003) in a study<br />

area of the Southern Alps, allowed the monitoring of spatiotemporal dynamics of forest,<br />

woodl<strong>and</strong> <strong>and</strong> tundra formations <strong>and</strong> pointed out significant upward shifts of their ecotones,<br />

formalised as forest, tree <strong>and</strong> tundra lines.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Quantitative assessment of temporal dynamics in altitudinal-driven ecotones<br />

355<br />

Analysis of l<strong>and</strong> cover dynamics by spatial overlay of multi-temporal l<strong>and</strong> cover maps showed<br />

that forest <strong>and</strong> scattered tree formations tend to exp<strong>and</strong> their areas, while tundra was more stable<br />

in time. However, the application of analyses addressing ecotone dynamics more specifically<br />

proved that the uppermost location of these formations underwent significant changes <strong>and</strong><br />

showed a general trend to advance in altitude.<br />

Ecotone dynamics typically correspond to a balance between colonization <strong>and</strong> extinctions,<br />

which in the case of mountain altitude-driven ecotones can be formalised as advances or retreats<br />

on the slopes. Therefore, to identify unambiguously if the changes correspond to an effective<br />

advance or retreat in altitude, we performed a directional analysis of the altitudinal shifts of<br />

extreme outposts along the maximum slope path. Results showed that most of the ecotone<br />

dynamics corresponded to advances or overpasses of outposts. <strong>Forest</strong> line retreats were virtually<br />

absent, corresponding at almost all outpost changes to advances in altitude. Tree <strong>and</strong> tundra<br />

lines showed a certain amount of retreats concentrated at the top of their current altitudinal<br />

distribution, but had a much higher rate of advance than retreat. The median values for the forest,<br />

tree <strong>and</strong> tundra line advances corresponded to decadal upward shifts of 26, 17 <strong>and</strong> 15 m<br />

respectively. Absolute altitudinal shift (considering together advances as positive values <strong>and</strong><br />

retreats as negatives in the same frequency distribution for each ecotone) had similar median<br />

values of 25, 13 <strong>and</strong> 11 m respectively.<br />

Spatial pattern analysis at the ecotone locations provided complementary information for a<br />

better underst<strong>and</strong>ing of their dynamics. Thus forest <strong>and</strong> tree lines tended to show a more<br />

compact arrangement in 2003 than in 1954, with a significant decrease in islets, indicating a<br />

decrease in long-distance dispersion <strong>and</strong> suggesting a stabilisation of its advance in the form of<br />

massive colonisation along wide advance fronts. Alternatively, the tundra line increased the<br />

number of islets, pointing towards an increase in long-distance colonisation. This fact, along<br />

with the intense dynamics deduced from the results of advance <strong>and</strong> retreat paths, points out this<br />

ecotone as prone to undergo significant future changes in the case of the continuing of present<br />

trends in environmental dynamics.<br />

References<br />

Calvo-Iglesias, M. S.; Fra-Paleo, U.; Crecente-Maseda, R. <strong>and</strong> Díaz-Varela, R. A., 2006.<br />

Directions of <strong>Change</strong> in L<strong>and</strong> Cover <strong>and</strong> L<strong>and</strong>scape Patterns from 1957 to 2000 in<br />

Agricultural <strong>L<strong>and</strong>scapes</strong> in NW Spain. Environmental Management, 38: 921-933.<br />

Cohen, J., 1960. A coefficient of agreement for nominal scales. Educational <strong>and</strong> Psychological<br />

Measurement, 20, 37–40<br />

Díaz-Varela, R. A.; Colombo, R.; Meroni, M.; Calvo-Iglesias, M. S.; Buffoni, A. <strong>and</strong><br />

Tagliaferri, A., 2010. Spatio-temporal analysis of alpine ecotones: A spatial explicit<br />

model targeting altitudinal vegetation shifts. Ecological Modelling, 221: 621-633.<br />

Didier, L., 2001. Invasion patterns of European larch <strong>and</strong> Swiss stone pine in subalpine pastures<br />

in the French Alps. <strong>Forest</strong> Ecology <strong>and</strong> Management, 145: 67-77.<br />

Körner, C., 1998. A re-assessment of high elevation treeline positions <strong>and</strong> their explanation.<br />

Oecologia 115, 445-459.<br />

Körner, C., 1999. Alpine Plant Life: Functional Plant Ecology of High Mountain Ecosystems.<br />

Springer, cop., Berlin, 330 pp.<br />

Theurillat, J.-P. <strong>and</strong> Guisan, A., 2001. Potential Impact of Climate <strong>Change</strong> on Vegetation in the<br />

European Alps: A Review. Climatic <strong>Change</strong>, 50: 77-109.<br />

Vogt, P., 2008. User guide of GUIDOS. Version 1.1. Institute for Environment <strong>and</strong><br />

Sustainability, European Commission, Joint Research Centre. Ispra (Italy).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Quantitative assessment of temporal dynamics in altitudinal-driven ecotones<br />

356<br />

Vogt, P.; Ferrari, J. R.; Lookingbill, T. R.; Gardner, R. H.; Riitters, K. H. <strong>and</strong> Ostapowicz, K.,<br />

2009. Mapping functional connectivity. Ecolgical Indicators, 9: 64-71<br />

Table 1. <strong>Change</strong> matrix between 1954 <strong>and</strong> 2003. Cell values are per-class areas (ha) resulting from the<br />

spatial overlay of 1954 <strong>and</strong> 2003 l<strong>and</strong> cover maps. Main diagonals correspond to stable areas while<br />

marginals correspond to changes. Per class Kappa indices are shown in the last column<br />

2003<br />

1954<br />

Dense forest Scattered trees Tundra Other Kappa<br />

Dense forest 259.84 101.04 19.12 203.04 0.43<br />

Scattered trees 0.96 87.76 89.72 26.20 0.42<br />

Tundra 1.08 17.16 1563.40 198.64 0.86<br />

Other 1.88 1.80 30.44 8979.28 0.98<br />

Figure 1: Study area<br />

Figure 2: Simplified representation of the conceptual approach for the definition of extreme outposts for a<br />

slope section<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


R.A. Diaz-Varela et al. 2010. Quantitative assessment of temporal dynamics in altitudinal-driven ecotones<br />

357<br />

Figure 3: Altitudinal distribution of the extreme outpost (forest line, tree line <strong>and</strong> tundra line ecotones) for<br />

the 1954 <strong>and</strong> 2003. Between brackets overall frequencies of raster map cells labelled as ecotones<br />

<strong>Forest</strong> line advance Tree line advance Tundra line advance<br />

<strong>Forest</strong> line retreat Tree line retreat Tundra line retreat<br />

Figure 4: Advances <strong>and</strong> retreats of the three ecotones in the period 1957-2003. In each graphic the X axis<br />

represents the initial altitude, Y axis the increment in horizontal distance <strong>and</strong> Z the increment in altitude<br />

(positive in advances <strong>and</strong> negative in retreats) for each of the events of altitudinal shifting<br />

80,00<br />

70,00<br />

60,00<br />

TREE54<br />

TREE03<br />

90,00<br />

80,00<br />

70,00<br />

TUNDRA54<br />

TUNDRAO3<br />

60,00<br />

50,00<br />

50,00<br />

40,00<br />

40,00<br />

30,00<br />

30,00<br />

20,00<br />

20,00<br />

10,00<br />

10,00<br />

0,00<br />

0,00<br />

core edge branch bridge islet loop<br />

core edge branch bridge islet loop<br />

Figure 5: Percentages of the different spatial pattern morphology classes for the targeted ecotone types<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Duarte & F.C. Rego 2010. L<strong>and</strong> use changes in Portugal between 1990 <strong>and</strong> 2005<br />

358<br />

L<strong>and</strong> use changes in Portugal between 1990 <strong>and</strong> 2005<br />

Inês Duarte 1,2* & Francisco Castro Rego 2<br />

1 Centro de Investigação em Ciências do Ambiente e Empresariais, INUAF, Convento<br />

Espírito Santo, 8100-641 Loulé, Portugal<br />

2 Centro de Ecologia Aplicada Prof. Baeta Neves, Tapada da Ajuda, 1300 Lisboa,<br />

Portugal<br />

Abstract<br />

In the last decades, changes occurred in all Mediterranean territory. Social <strong>and</strong> economical<br />

trends changed <strong>and</strong> the related l<strong>and</strong> use accompanied this alteration. We used different<br />

Portuguese cartography (1990 <strong>and</strong> 2005), made them comparable <strong>and</strong> identified the changes on<br />

the territory. Despites all the incentives for forestation in the last decades, results showed that<br />

shrubl<strong>and</strong>s advanced on the territory. Eucalyptus forests <strong>and</strong> irrigated agriculture increased too<br />

with less significance. The general trends verified were for shrubl<strong>and</strong>s increase <strong>and</strong> lost of<br />

traditional uses.<br />

Keywords: Portugal, L<strong>and</strong> use changes, L<strong>and</strong> use types, <strong>Forest</strong>s, Agriculture, Shrubl<strong>and</strong>s,<br />

Oaks, Transition<br />

1. Introduction<br />

The l<strong>and</strong>scape is a reflection of the aims of the society. Neolithic man began changing the<br />

l<strong>and</strong>scape through the invention of agriculture (Pasquale et al, 2004). In XIX century he made<br />

long railways <strong>and</strong> enlarged constructed surfaces. Portugal, like all Mediterranean countries,<br />

have a l<strong>and</strong>scape with centuries of history from intervention, cutting, grazing, fire, <strong>and</strong> the<br />

socio-economical trends always were related with agriculture, forestry <strong>and</strong> livestock (Duarte et<br />

al, 2009; Acácio, 2009; Lloret, 2003). The traditional l<strong>and</strong> use has last for centuries, or<br />

millennia, but recently major changes have been occurring (Paqualle et al, 2004). Many<br />

agricultural l<strong>and</strong>s were ab<strong>and</strong>oned, due to migration of people to urban areas in search of better<br />

socio-economic conditions. The silvicultural practices, related with the traditional economie, as<br />

also decreased largely. Grazing became negligible. All these events determined the increase in<br />

shrubl<strong>and</strong> areas <strong>and</strong> forest cover, continue, reducing the fragmentation of the l<strong>and</strong>scape<br />

(Paqualle et al, 2004; Hill et al, 2004; Regato-Pajares et al, 2004).<br />

In the first half of the XX century, in Portugal, the agriculture exp<strong>and</strong>s with the campaign of<br />

wheat (Duarte et al, 2009). Then, in the second half of the century, the forests take a lot of his<br />

place <strong>and</strong> shrubl<strong>and</strong>s increased because of the ab<strong>and</strong>onment of the l<strong>and</strong> (Ferreira el al, 2004).<br />

More recently, agricultural l<strong>and</strong> increased again with the irrigation possibilities <strong>and</strong><br />

diversification of productions (Pinto Correia et al, 2006; Ferreira et al, 2004). But forests still<br />

the more representative l<strong>and</strong> use types (Ferreira et al, 2004).<br />

There has been some studies about changes in l<strong>and</strong> use in the Portuguese territory (Ferreira at<br />

all, 2004; Pinto Correia et al, 2006) that shows this occurrences in the portuguese context.<br />

* Corresponding author. Tel.:+351 961 947 466 - Fax:+351 289 420 478<br />

Email address: inesmarquesduarte@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Duarte & F.C. Rego 2010. L<strong>and</strong> use changes in Portugal between 1990 <strong>and</strong> 2005<br />

359<br />

However, with the XXI century arrival, it matters to underst<strong>and</strong> if the trends are maintaining or,<br />

if something changed in the l<strong>and</strong>-cover relation with society.<br />

With this study, we pretend to identify (1) the main changes in l<strong>and</strong> use in Portugal between<br />

1990 <strong>and</strong> 2005, (2) the general trends <strong>and</strong> (3) the dynamic relationship between the l<strong>and</strong> use<br />

types.<br />

2. Methodology<br />

We used as a basis for study two cartografic elements: Carta de Ocupação de Solo -1990<br />

(COS90) from the responsibility of the Portuguese Geographic Institute (IGP) <strong>and</strong> the National<br />

<strong>Forest</strong> Inventory 2005 (IFN05) of the National <strong>Forest</strong> Authority (Ministry of Agriculture <strong>and</strong><br />

Fisheries). However, these mappings have been developed with different objectives. The first,<br />

inspired in the nomenclature of CORINE l<strong>and</strong>cover, aims to give greater accuracy <strong>and</strong> detail to<br />

the Corine coverage, in Portugal (Caetano et al, 2007), assumed at the time as a tool to support<br />

planning studies, <strong>and</strong> decision making. IFN05 was conceived for supporting policies <strong>and</strong> forest<br />

management (DGRF, 2007).<br />

Thus, COS90 is presented on the form of polygons <strong>and</strong> is based on a alphanumeric information<br />

legend as type of occupation, first <strong>and</strong> second dominant species <strong>and</strong> a reference to density. In<br />

this map we obtain the total of 638 different l<strong>and</strong> use type combinations. The IFN05, is<br />

presented on the form of 334390 points, in a grid of 2km for 2km, covering all continental<br />

Portuguese territory. This legend, also alphanumeric, but with a different nomenclature, was<br />

established for purposes of identification of type of use, dominant species, second <strong>and</strong> third<br />

dominant species, vertical structure, density <strong>and</strong> presence of other uses. Overall, we obtain 2163<br />

l<strong>and</strong> use type combinations.<br />

Methodology begans on data transformation in order to compare both cartografic elements. We<br />

used ArcGis 9.3 program package, to join tables, getting a grid of points with the information<br />

contained in each of the mappings in accordance with their original captions.<br />

Then we reclassified all the 334390 points, covering the whole territory, for the year 1990 <strong>and</strong><br />

for the year 2005, with a general common nomenclature. This new classification was based, on<br />

the first level, on the type of l<strong>and</strong> use: <strong>Forest</strong>, agriculture, shrubl<strong>and</strong>s, other (artificial) <strong>and</strong><br />

wetl<strong>and</strong>s. On the second level, considered the dominant specie (forest) or type of agriculture. In<br />

the class shrubl<strong>and</strong>s were included also clear cut, burnt, sown <strong>and</strong> young plantation areas as<br />

well as recently soil mobilized. We removed wetl<strong>and</strong>s <strong>and</strong> remaining 328029 points.<br />

The comparability has made possible, such as the construction of the transition matrix, using the<br />

Microsoft Office Excel 2007.<br />

3. Results<br />

To identify the l<strong>and</strong> use changes between 1990 <strong>and</strong> 2005, we produced a transition matrix that<br />

explains the alterations between classes, showed in table 1. With this matrix, <strong>and</strong> figure 1<br />

support, we can clearly verify the (1) significant increasing of shrubl<strong>and</strong>s, (2) the visible<br />

increase of Eucalyptus, irrigated agriculture <strong>and</strong> artificial occupancy, <strong>and</strong> (3) the losses of Pine,<br />

Cork & Holm oaks <strong>and</strong> unirrigated <strong>and</strong> other agricultures.<br />

The persistence index calculated was only 51%. This means that 49% of Portuguese territory<br />

suffered a transition for one other class, during the 15 years in analysis.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Duarte & F.C. Rego 2010. L<strong>and</strong> use changes in Portugal between 1990 <strong>and</strong> 2005<br />

360<br />

Table 1: Transition matrix between 1990 <strong>and</strong> 2005<br />

Shrub Artificial Cork<br />

&<br />

Other<br />

forest<br />

Eucalyptus Pine Oaks Irrigated<br />

agriculture<br />

Unirrigated<br />

agriculture<br />

Other<br />

agriculture<br />

Total<br />

2005<br />

Holm<br />

oaks<br />

sp<br />

Shrubl<strong>and</strong>s 31052 3544 7112 2497 2561 16540 2501 728 9876 9597 86008<br />

Artificial 1097 7502 251 240 428 1273 67 481 1860 2172 15371<br />

Cork &<br />

846 45 26247 775 465 1892 289 85 2045 1671 34360<br />

Holm oaks<br />

Other forest 968 133 1179 2313 278 1323 436 257 1284 1214 9385<br />

sp<br />

Eucalyptus 2296 346 645 524 15144 7874 118 265 1353 1012 29577<br />

Pine 3708 278 514 1482 2014 19927 478 254 1376 1151 31182<br />

Oaks 986 54 551 655 48 1042 1563 104 703 493 6199<br />

Irrigated 315 359 1228 163 130 273 21 5850 10486 2807 21632<br />

agriculture<br />

Unirrigated 2692 864 5759 548 359 1108 347 4171 32302 9938 58088<br />

agriculture<br />

Other<br />

1827 1055 1234 393 229 1161 230 1298 9441 19359 36227<br />

agriculture<br />

Total 1990 45787 14180 44720 9590 21656 52413 6050 13493 70726 49414 328029<br />

Analyzing the data, in figure 2, it is perceptible the gains <strong>and</strong> losses from each l<strong>and</strong> use type<br />

from <strong>and</strong> to each other. So, shrubl<strong>and</strong>s, including clear cut, burnt, sown <strong>and</strong> young plantation<br />

areas, were the most significant transition report. The gains came mostly from Pine, unirrigated<br />

<strong>and</strong> other agricultures. The losses, with low proportions, were curiously with the same classes<br />

<strong>and</strong> with Eucalyptus. The exchanges with Cork & Holm were significant too, with just one way<br />

to shrubl<strong>and</strong>s <strong>and</strong> no return.<br />

100000<br />

L<strong>and</strong> use variation between 1990 <strong>and</strong> 2005<br />

90000<br />

80000<br />

70000<br />

Nr. points<br />

60000<br />

50000<br />

40000<br />

30000<br />

20000<br />

10000<br />

0<br />

Shrubl<strong>and</strong>s<br />

Artificial<br />

Cork &<br />

Holm oaks<br />

Other<br />

forest sp<br />

Eucalyptus Pine Oaks<br />

Irrigated<br />

agriculture<br />

Unirrigated<br />

agriculture<br />

Other<br />

agriculture<br />

1990 46207 14368 44513 9654 21670 52448 6063 13615 71357 49573<br />

2005 86325 15499 34429 9512 29605 31240 6216 21816 58196 36264<br />

Figure 1: Comparison, for each l<strong>and</strong> use type, between 1990 <strong>and</strong> 2005<br />

Also in figure 2, despite Artificial l<strong>and</strong> cover, mostly urban areas, had a slightly increase, the<br />

transitions with it were significant. The transitions from this one to shrubl<strong>and</strong>s were notable.<br />

This is probably a sign of ab<strong>and</strong>onment. In the other way, a lot of agricultural <strong>and</strong> Pine forests<br />

switches to artificial areas.<br />

Cork & Holm oak forests decreased, turning into shrubl<strong>and</strong>s or unirrigated agriculture. The first<br />

one, due to ab<strong>and</strong>onment, the second, probably correlated with the mortality of Cork & Holm,<br />

leaving only the agricultural part of the system.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Duarte & F.C. Rego 2010. L<strong>and</strong> use changes in Portugal between 1990 <strong>and</strong> 2005<br />

361<br />

Shrubl<strong>and</strong>s<br />

Artificial<br />

20000<br />

3000<br />

15000<br />

2000<br />

1000<br />

10000<br />

To Shrublans<br />

0<br />

To artificial<br />

5000<br />

From Shrublans<br />

‐1000<br />

From Artificial<br />

0<br />

‐2000<br />

‐3000<br />

‐5000<br />

‐4000<br />

Eucalyptus forests<br />

Pine forests<br />

10000<br />

5000<br />

8000<br />

0<br />

6000<br />

4000<br />

To Eucalyptus<br />

‐5000<br />

To Pine<br />

2000<br />

From Eucalyptus<br />

‐10000<br />

From Pine<br />

0<br />

‐2000<br />

‐15000<br />

‐4000<br />

‐20000<br />

Cork & Holm oaks<br />

Other oaks forests<br />

4000<br />

2000<br />

2000<br />

1500<br />

1000<br />

0<br />

‐2000<br />

‐4000<br />

To Cork & Holm oaks<br />

From Cork & Holm oaks<br />

500<br />

0<br />

‐500<br />

‐1000<br />

‐1500<br />

To Other Oaks<br />

From Other Oaks<br />

‐6000<br />

‐2000<br />

‐2500<br />

‐8000<br />

‐3000<br />

Other forests species<br />

Irrigated agriculture<br />

2000<br />

12000<br />

1500<br />

10000<br />

1000<br />

8000<br />

500<br />

0<br />

‐500<br />

‐1000<br />

‐1500<br />

To Other forests<br />

From Other forests<br />

6000<br />

4000<br />

2000<br />

0<br />

To Irrtigated agriculture<br />

From Irrigated agriculture<br />

‐2000<br />

‐2000<br />

‐2500<br />

‐4000<br />

‐3000<br />

‐6000<br />

Other Agriculture<br />

Unirrigated agriculture<br />

15000<br />

15000<br />

10000<br />

10000<br />

5000<br />

5000<br />

0<br />

To Other agriculture<br />

From Other agriculture<br />

0<br />

To Unirrigated agriculture<br />

From Unirrigated agriculture<br />

‐5000<br />

‐5000<br />

‐10000<br />

‐10000<br />

‐15000<br />

‐15000<br />

Figure 2: Representative graphs of gains <strong>and</strong> losses in each l<strong>and</strong> use type, related with each one other<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Duarte & F.C. Rego 2010. L<strong>and</strong> use changes in Portugal between 1990 <strong>and</strong> 2005<br />

362<br />

Many Eucalyptus areas turned over into Pine forests. Furthermore, areas of Pine forests were<br />

converted into Eucalyptus, or degraded forests (shrubl<strong>and</strong>s). This can be justified by the forest<br />

fires that had been occur frequently over Pine forest in Portugal during this years (Acácio,<br />

2009).<br />

Has showed in figure 2 other oak forests are being turned into shrubl<strong>and</strong>s (ab<strong>and</strong>onment or fire)<br />

or into Pine forests. The last case can be justified with the introduction of Pine into oak systems,<br />

gettin dominance over the years.<br />

Between the agricultural types of l<strong>and</strong> uses, switches occured, with irrigated agriculture areas<br />

growing, <strong>and</strong> a part of the others loosing areas to shrublans once again.<br />

Figure 3 represents the dinamic changes between the studied l<strong>and</strong> uses. On this figure were<br />

considered only transitions with more meaningful values, i.e. greater than 10% of each class or<br />

higher than 3000 points. The only exceptions considered were the transitions from shrubl<strong>and</strong>s to<br />

Pine forest or Eucalyptus forest because they are significant on these forest systems dynamics.<br />

It is clear to identify a group of agricultural uses <strong>and</strong> one other of forest uses. Cork & Holm had<br />

caracters from both, staying undistiguished from them. On this representation it is notable the<br />

magnetic effect of Shul<strong>and</strong>s, that justifies the values obtained on table 1 <strong>and</strong> figure 1.<br />

On the agricultural part of studied l<strong>and</strong> uses, many changes occured, maybe related with<br />

rotational l<strong>and</strong> programs, between irrigated, unirrigated <strong>and</strong> other types of agriculture. The last<br />

two, with frequent losses for the class of shrubl<strong>and</strong>s.<br />

Figure 3: Representation of the dynamic changes of l<strong>and</strong> use<br />

Artificial areas losses were significant to shrubl<strong>and</strong>s, however, as shown in figures 1 <strong>and</strong> 2, less<br />

significant gains of other l<strong>and</strong> uses justified an increase in the total value of artificial areas.<br />

4. Discussion<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Duarte & F.C. Rego 2010. L<strong>and</strong> use changes in Portugal between 1990 <strong>and</strong> 2005<br />

363<br />

On this study, we analyze the changes on soil occupancy in continental Portugal. The index of<br />

persistence obtained was 51%. Pinto Correia et al (2006), calculated between 1990 <strong>and</strong> 2000,<br />

for the same territory, a persistence index of 86,7 %. It means that in the last five years, changes<br />

were accelerated.<br />

The general trends, despite all the incentives for forestation in recent decades, were for the<br />

decrease of forest area <strong>and</strong> increase of shrubl<strong>and</strong>s areas. Eucalyptus <strong>and</strong> irrigated agriculture<br />

also increased. Also occurred one slight increase in artificial areas, such as in other agriculture.<br />

The dynamical model identified let us concerned about the trends in portuguese <strong>and</strong> all<br />

mediterranean territory, because of the forest degradation <strong>and</strong> losses of natural values as corks<br />

forests <strong>and</strong> unirrigated agriculture. On the other h<strong>and</strong>, perhaps this is a time of change, of<br />

reflection, similar to the socio-economic reflexion of the day.<br />

References<br />

Acácio, V. 2009. The dinamics of cork oak systems in portugal: the role of ecological <strong>and</strong> l<strong>and</strong><br />

use factors. Thesis. Wageningen, Netherl<strong>and</strong>. Wageningen University: 207 p.<br />

Caetano, M., V. Nunes <strong>and</strong> M. Pereira, 2009. L<strong>and</strong> use <strong>and</strong> l<strong>and</strong> cover map of continental<br />

Portugal for 2007 (COS2007): project presentation <strong>and</strong> technical specifications<br />

development, 3rd Workshop of the EARSeL Special Interest Group Remote Sensing of<br />

L<strong>and</strong> Use <strong>and</strong> L<strong>and</strong> Cover, Bonn, Germany<br />

DGRF. 2007. Resultados do Inventário Florestal Nacional 2005-2006. Direcção Geral dos<br />

Recursos Florestais. Lisboa<br />

Duarte, I. Rego, F. Fonseca, L. 2009. Avaliação da Regeneração da Paisagem após o incêndio<br />

de 2004 na Serra do Caldeirão.Coimbra. Portugal. APDR. Revista Portuguesa de Estudos<br />

Regionais (20): 41-60<br />

Ferreira, P; Almeida, M; Fern<strong>and</strong>es, A; Codipietro, P; Rego, F. 2004. L<strong>and</strong>scape Dynamicsin<br />

the Area of Serra da Arrábida <strong>and</strong> the Sado River Estuary. In Mazzoneli, S; di Pasquale,<br />

G; Mulligan, M; Martino, P; Rego, F. Recent Dynamics of the Mediterranean Vegetation<br />

<strong>and</strong> Lanscape. Engl<strong>and</strong>. John Wiley & Sons:201-209<br />

Lloret, F; Vilà, M. 2003. Diversity patterns of plant functional types in relation to fire regime<br />

<strong>and</strong> previos l<strong>and</strong> use in mediterranean woodl<strong>and</strong>s. Ludvike. Sweeden. Opulus Press<br />

Uppsala. Journal of Vegetation Science (14): 387-398<br />

Pasquale, G; Martino, P; Mazzoleni, S. 2004. <strong>Forest</strong> History in the Mediterranean Region. In:<br />

Mazzoneli, S; di Pasquale, G; Mulligan, M; Martino, P; Rego, F. Recent Dynamics of the<br />

Mediterranean Vegetation <strong>and</strong> Lanscape. Engl<strong>and</strong>. John Wiley & Sons:13-20<br />

Pinto Correia, T; Breman, B; Jorge, V; Dneboská, M. 2006. Estudo sobre o ab<strong>and</strong>ono em<br />

Portugal Continental, Análise das dinâmicas da ocupação do solo, do sector agrícola e<br />

da comunidade rural, tipologia de áreas rurais. Portugal. Universidade de Évora. 201p.<br />

Quézel, P. 2004. Large-scale Post Glacial Distribution of Vegetation Structures in<br />

Maditerranean Region. In: Mazzoneli, S; di Pasquale, G; Mulligan, M; Martino, P; Rego,<br />

F. Recent Dynamics of the Mediterranean Vegetation <strong>and</strong> Lanscape. Engl<strong>and</strong>, John<br />

Wiley & Sons: 3-12<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.S.V. Filho & M.L. Leite 2010. Simulation of climate scenarios for the region of Campos Gerais, State of Parana, Brazil<br />

364<br />

Simulation of climate scenarios for the region of Campos Gerais, State<br />

of Parana, Brazil<br />

Jorim Sousa das Virgens Filho 1* & Maysa de Lima Leite 2<br />

1<br />

Departamento de Matemática e Estatística, Universidade Estadual de Ponta Grossa,<br />

Brazil<br />

2<br />

Departamento de Biologia Geral, Universidade Estadual de Ponta Grossa, Brazil<br />

Abstract<br />

This study aimed to simulate climate scenarios based on possible change to the region of<br />

Campos Gerais, state of Parana, Brazil. Originally defined as a phytogeographical region, the<br />

Campos Gerais underst<strong>and</strong> the grassl<strong>and</strong>s <strong>and</strong> savanna parks situated on the edge of the Second<br />

Paraná Plateau. In the forests of Campos Gerais, the Araucaria angustifolia is the main tree<br />

species, occupying portions of the plateau state of Parana whose floristic composition is<br />

strongly influenced by low temperatures <strong>and</strong> frost occurrence. Thus, using daily weather series,<br />

stochastic climate models were parameterized to simulate the climate scenarios, based on<br />

projections of the IPCC. The results achieved through analysis of graphs, presenting essential<br />

elements for a systematic reflection on the future of the floristic diversity of Campos Gerais,<br />

showed that an environment in the near future may be unfavorable to the development of<br />

species that today fully supplies the forests in this region.<br />

Keywords: simulation, climate scenarios, temperature, precipitation, Araucaria angustifolia.<br />

1. Introduction<br />

The Earth has been undergoing continuous natural oscillations along its evolution, creating new<br />

organizations <strong>and</strong> changing ecosystems. Among the components that operate in this dynamic<br />

system in itself, the weather is a factor that interferes directly or indirectly in these<br />

transformations, causing instability in the natural l<strong>and</strong> surface <strong>and</strong> oceans. Thus, as<br />

recommended by Claussen in 2004, it is expected that an important aspect of global climate<br />

change, is their likely impact on natural ecosystems considering that climate influences soil <strong>and</strong><br />

biota in the same way that it can also be strongly modulated by biological processes of<br />

vegetation, phytoplankton <strong>and</strong> other characteristics of the biosphere. Within a bioclimatological<br />

perspective, according to Worbes 2002, in tropical forests, seasonal precipitation regimes or<br />

flooding is described as the major determinants of the seasonality of growth. However, in<br />

subtropical forests, the annual variation of air temperature may have great influence on the<br />

regulation of cambial activity. Originally defined as a phytogeographical region, the Campos<br />

Gerais of Parana underst<strong>and</strong> grassl<strong>and</strong>s <strong>and</strong> forests galleries or geldings isolates, mixed in a<br />

temperate rainforest, located on the edge of the Second Parana Plateau (Maack 1948). This<br />

forest type can be defined as a mixing of floras from different sources, with typical<br />

phytophysiognomic patterns in a characteristically rainy climatic zone, without direct influence<br />

of the ocean, but with well distributed precipitation throughout the year. According to Roderjan<br />

et al. 2002, forests in Campos Gerais, the Araucaria or Pinheiro-do-Paraná (Araucaria<br />

angustifolia (Bertol.) Kuntze) is the main tree species, occupying portions of the State of Parana<br />

Plateau whose floristic composition is strongly influenced by low temperatures <strong>and</strong> the<br />

* Corresponding author. Tel.: +55 (42) 3220 3050<br />

Email address: jvirgens@uepg.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.S.V. Filho & M.L. Leite 2010. Simulation of climate scenarios for the region of Campos Gerais, State of Parana, Brazil<br />

365<br />

occurrence of regular frosts in winter. The Araucaria occurs in regions with annual average<br />

temperatures ranging from 12°C to 18°C, supporting frost until -10°C, characterized therefore<br />

as a species of temperate climate. The climate of the region of Campos Gerais, Parana,<br />

according to Köppen may occur variably in some places, from subtropical (Cfa) to temperate<br />

(Cfb), with an average temperature of the coldest month below 18°C <strong>and</strong> average temperature in<br />

warmest month above 22°C, with frequent frosts in winter <strong>and</strong> trend of concentration of<br />

precipitation during the summer months, but without a dry season (IAPAR 2000). According to<br />

Salazar et. al 2007, the geographic distribution of vegetation <strong>and</strong> its relationship to climate, has<br />

been object of analysis through models of biomes or biogeographic models, which in turn use it<br />

as the central paradigm assumption that climate has a dominant control on distribution of<br />

vegetation, simulating potential vegetation, based on some climatic parameters like temperature<br />

<strong>and</strong> precipitation. In this context, on the premise that the crops may be considered relevant<br />

probabilistic variables <strong>and</strong> depend on climatic factors during the growing season, it becomes<br />

important to the development of simulation models that generate synthetic data of climate, in<br />

order to reproduce, probabilistically, the occurrence of climatic components. Several climatic<br />

data generators are cited in the literature, in order to simulate sets of climate data, <strong>and</strong> create<br />

climate scenarios in order to provide alternative methods for measuring the risk of climate<br />

uncertainty that is related to alternative managements of enterprises conducted in<br />

agroecosystems (Virgens Filho 2001). Given the climatic transition projected by the<br />

Intergovernmental Panel on Climate <strong>Change</strong> (IPCC) that will affect natural resources,<br />

economics <strong>and</strong> societies around the world, actually unknown in magnitude, this study aimed to<br />

simulate climate scenarios based on possible climate change in the Region of Campos Gerais,<br />

Parana State, Brazil.<br />

2. Methodology<br />

This research was conducted at the State University of Ponta Grossa (UEPG), Parana State,<br />

Southern Brazil, where there were used historical series of climatic data of temperature <strong>and</strong><br />

precipitation, obtained from the Instituto Agronômico do Paraná (IAPAR) <strong>and</strong> the Instituto<br />

Nacional de Meteorologia (INMET) as shown in Table 1 for four cities in the region of Campos<br />

Gerais, Parana State (Figure 1). For the simulation of climate scenarios, it was used a beta<br />

version of the Stochastic Generator of Climatic Scenarios PGECLIMA_R, developed by the<br />

Computational <strong>and</strong> Applied Statistics Laboratory, in UEPG. The scenarios generated for the air<br />

temperature were based on climate projections for 2100, published in 2001 in the Third<br />

Assessment Report of the IPCC, in which its worst scenario is estimated to increase average<br />

global temperature of 5.8ºC <strong>and</strong> at its best scenario, an average increase of 1.4°C (IPCC 2001).<br />

The scenarios for the simulation of precipitation, were created from the variation of 10% for<br />

each degree Celsius in 2100, as suggested by Pike in 2005. Therefore, as the scenarios of air<br />

temperature were simulated for an increase of 1.4°C <strong>and</strong> 5.8°C, the average increase in<br />

precipitation was 14% <strong>and</strong> 58% in the total annual in 2100, for the worst <strong>and</strong> best climatic<br />

scenario, respectively. The analysis <strong>and</strong> discussion of the results were based on graphics, which<br />

projected the trend of the next 99 years (2002-2100), for simulated scenarios for temperature<br />

<strong>and</strong> precipitation.<br />

3. Result<br />

Figure 2 shows the graphics with the trends of annual mean air temperature for the four cities<br />

located in the Campos Gerais region of Parana, <strong>and</strong> for each site are presented separately, the<br />

worst <strong>and</strong> best simulated scenarios. It is observed by the simulated trends (gray line), that for<br />

the city of Telêmaco Borba, which has an average annual temperature of 19.5°C in the worst<br />

scenario, the projected average temperature for 2052 was approximately 21.5°C, while close to<br />

year 2100 it is projected to be around 25.6°C. In the best scenario, the temperature values in<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.S.V. Filho & M.L. Leite 2010. Simulation of climate scenarios for the region of Campos Gerais, State of Parana, Brazil<br />

366<br />

2052 <strong>and</strong> 2100 were approximately 20.1°C <strong>and</strong> 21°C, respectively. For the city of Castro, with<br />

an average annual temperature of 18°C, in the worst scenario for 2052, simulated temperature<br />

was 20°C, while close to year 2100, it was around 23.9°C. In the best scenario, the approached<br />

temperature for the years 2052 <strong>and</strong> 2100 were 18.5°C <strong>and</strong> 19.5°C, respectively. Note by the<br />

simulated trends, that for the city of Ponta Grossa, with an average annual temperature of<br />

18.7°C , in the worst scenario, the average temperature expected for 2052 was approximately<br />

20.5°C, while close to the year 2100 to it was projected to be around 24.6°C. In its best scenario<br />

the temperature in 2052 <strong>and</strong> 2100 will be close of 19.1°C <strong>and</strong> 20°C, respectively.<br />

Table 1 - Geographical coordinates of the cities <strong>and</strong> information related to the climatological series.<br />

Municipal<br />

District<br />

Latitude Longitude Elevation (m) Period<br />

Telêmaco Borba - 24°20’ - 50°37’ 768.0<br />

1972-<br />

2001<br />

Castro - 24°47’ - 50°00’ 1008.8 1972-<br />

Ponta Grossa - 25°13’ - 50°01’ 880.0 1972- 2001<br />

Lapa - 25°47’ - 49°46’ 910.0 1972- 2001<br />

2001<br />

Data<br />

Source<br />

IAPAR<br />

INMET<br />

IAPAR<br />

IAPAR<br />

Figure 1 - Location of the cities in the Region of Campos Gerais, Parana State, Brazil.<br />

For the city of Lapa, which average annual temperature is 17.9°C, in the worst scenario the<br />

average temperature generated for 2052 was 19.8°C, while in the year 2100 it was around<br />

23.8°C. In the best scenario, the temperature approached 18.3°C <strong>and</strong> 19.2°C for the years 2052<br />

<strong>and</strong> 2100, respectively. Figure 3 shows the graphics with the trends of annual totals of<br />

precipitation for the four cities in the region of Campos Gerais, Parana, <strong>and</strong> for each site are<br />

presented separately, the worst <strong>and</strong> best simulated scenarios. It was found by the simulated<br />

trends (gray line), that for the city of Telêmaco Borba, which has a historical average annual<br />

total precipitation of 1630 mm, that for the worst scenario, the projected annual total for 2052,<br />

was approximately 2049 mm, while next to the year 2100 it was projected to be around 2541<br />

mm. In its best scenario, the annual total in 2052 <strong>and</strong> 2100 was around 1571 mm <strong>and</strong> 1880 mm,<br />

respectively. For the city of Castro, with a historical average annual total precipitation of 1414<br />

mm, in the worst scenario, the annual total raised to 1922 mm in 2052, while in the year of<br />

2100 it was around 2273 mm. In the best scenario, annual total in 2052 <strong>and</strong> 2100 approached to<br />

1602 mm <strong>and</strong> 1710 mm, respectively. Note by the simulated trends, that for the city of Ponta<br />

Grossa, which has an annual historical average total of 1613 mm, in the worst scenario, the<br />

expected annual total for 2052 was approximately 2002 mm, while next to the year 2100, it was<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.S.V. Filho & M.L. Leite 2010. Simulation of climate scenarios for the region of Campos Gerais, State of Parana, Brazil<br />

367<br />

around 2493 mm. In its best scenario, the annual total in 2052 <strong>and</strong> 2100 was around 1776 mm<br />

<strong>and</strong> 1882 mm, respectively. For the city of Lapa, where the annual historical average total was<br />

1651 mm, in the worst simulated scenario, the annual total for 2052 was 1932 mm, while in the<br />

year of 2100, it approached 2246 mm. In the best scenario, the annual total in 2052 <strong>and</strong> 2100<br />

approached, 1911 mm <strong>and</strong> 1955 mm, respectively.<br />

Figure 2 - Mean air temperature scenarios simulated by PGECLIMA_R for the year 2100, considering the<br />

best <strong>and</strong> worst outlook projected by the IPCC.<br />

4. Discussion<br />

Given the fact that the region of Campos Gerais of Parana is located in a climatic zone,<br />

dominated in some places by the subtropical climate (Cfa) <strong>and</strong> others by temperate climate<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.S.V. Filho & M.L. Leite 2010. Simulation of climate scenarios for the region of Campos Gerais, State of Parana, Brazil<br />

368<br />

(Cfb), with precipitation concentrated during summer months, without dry season <strong>and</strong> with<br />

regular occurrences of frost in winter, it is expected that from a perspective of regional climate<br />

change, the increase in temperature causes significant changes in the l<strong>and</strong>scape since the<br />

floristic composition of the forest <strong>and</strong> grassl<strong>and</strong> are strongly influenced by temperature <strong>and</strong><br />

precipitation.<br />

Figure 3 - Precipitation scenarios simulated by PGECLIMA_R for the year 2100, considering the best <strong>and</strong><br />

worst outlook projected by the IPCC.<br />

In the case of Araucaria <strong>Forest</strong>, which is a forest formation adapted to conditions of humid<br />

altitude, its main plant species, the Araucaria, have their reproductive phenology affected<br />

mainly by abiotic factors such as temperature <strong>and</strong> photoperiod (Liebsch <strong>and</strong> Mikich 2009).<br />

According to Zanon 2007, the increase in temperature <strong>and</strong> precipitation have a positive<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.S.V. Filho & M.L. Leite 2010. Simulation of climate scenarios for the region of Campos Gerais, State of Parana, Brazil<br />

369<br />

influence on the annual growth rings, but the occurrence of precipitation coupled with low<br />

temperatures reduces the growth of them. With the simulated climate scenarios, presented in<br />

this research, for the four cities of Campos Gerais, Parana State, one can deduce that at best,<br />

which projected an average increase of 1.4°C in the average temperature <strong>and</strong> 14% increase in<br />

precipitation, future changes will be possibly positive as regards the development of Araucaria<br />

<strong>Forest</strong>. However, for a scenario with an increase of 5.8°C in the average temperature <strong>and</strong> 58%<br />

increase in the amount of precipitation, extreme events may occur daily for these climatic<br />

elements which, according to Kramer <strong>and</strong> Kozlowski 1960, influence negatively the hydration<br />

<strong>and</strong> dehydration of the trees, which may result in variations in the diameter growth during the<br />

twenty-four hours a day.<br />

Acknowlegments<br />

The authors thank to the Instituto Agronômico do Paraná (IAPAR) <strong>and</strong> Instituto Nacional de<br />

Meteorologia (INMET) for the authorization to access the climatic data, <strong>and</strong> to the CNPq e<br />

Fundação Araucária, by the research support.<br />

References<br />

Claussen, M., 2004. Does l<strong>and</strong> surface matter in weather <strong>and</strong> climate In: Kabat, P., Claussen,<br />

M., Dirmeyer, P.A., Gash, J.H.C., Bravo de Guenni, L., Meybeck, M., Pielke Sr, R.A.,<br />

Vörösmarty, C.J., Hutjes, R.W.A. <strong>and</strong> Lütkemeier, S. (Eds.). Vegetation, water, humans<br />

<strong>and</strong> climate: a new perspective on an interactive system. IGBP <strong>Global</strong> <strong>Change</strong> Series. New<br />

York, Springer-Verlag: part A, p. 5-6.<br />

IAPAR - Instituto Agronômico do Paraná, 2000. Cartas climáticas do Estado do Paraná.<br />

IAPAR, Londrina, Brasil, CD-ROM Versão 1.0.<br />

Intergovernmental Panel on Climate <strong>Change</strong> – IPCC, 2001. Climate <strong>Change</strong> 2001: The<br />

Scientific Basis-Contribution of Working Group 1 to the IPCC Third Assessment Report.<br />

Cambridge Univ. Press, Cambridge, UK, 17 p.<br />

Kramer, P.J. <strong>and</strong> Kozlowski, T., 1960. Fisiologia das árvores. Fundação Caloustre. Gulbenkian<br />

Lisboa, Portugal, 745 p.<br />

Liebsch, D. <strong>and</strong> Mikich, S.B., 2009. Fenologia reprodutiva de espécies vegetais da Floresta<br />

Ombrófila Mista do Paraná, Brasil. Revista Brasileira de Botânica, 32: 375-391.<br />

Maack, R., 1948. Notas preliminares sobre clima, solos e vegetação do Estado do Paraná.<br />

Arquivos de Biologia e Tecnologia, 2: 102-200.<br />

Pyke, C.R., 2005. Interactions between habitat loss <strong>and</strong> climate change: implications for fairy<br />

shrimp in the central valley ecoregion of california, USA. Climatic <strong>Change</strong>, 68: 199–218.<br />

Roderjan, C.V., Galvão, F., Kuniyoshi, Y.S. <strong>and</strong> Hatschbach, G.G., 2002. As unidades<br />

fitogeográficas do Estado do Paraná, Brasil. Ciência & Ambiente, 24: 78-118.<br />

Worbes, M., 2002. One hundred years of tree-ring research in the tropics - a brief history <strong>and</strong> an<br />

outlook to future challenges. Dendrochronologia, 20: 217-231.<br />

Virgens Filho, J. S., 2001. Ferramenta computacional para simulação de séries climáticas<br />

diárias, baseada na parametrização dinâmica das distribuições de probabilidade. Tese<br />

(Doutorado em Agronomia/Energia na Agricultura) – UNESP, Botucatu-SP, Brasil, 92 p.<br />

Zanon, M.L.B., 2007. Crescimento da Araucaria angustifolia (Bertol.) Kuntze. diferenciado por<br />

dioicia. Tese (Doutorado em Engenharia Florestal) – UFSM, Santa Maria-RS, Brasil, 110 p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F.J. Gómez et al. 2010. L<strong>and</strong> use changes <strong>and</strong> mixed forest dynamics. The case of Montiferru Mountains<br />

370<br />

L<strong>and</strong> use changes <strong>and</strong> mixed forest dynamics. The case of Montiferru<br />

Mountains (Sardinia, Italy) (1955-2006)<br />

Francisco Javier Gómez * , Martí Boada & Diego Varga<br />

Institute of Environmental Science <strong>and</strong> Technology, Autonomous University of<br />

Barcelona (ICTA-UAB), Spain<br />

Abstract<br />

L<strong>and</strong> cover dynamics in forest l<strong>and</strong>scapes are altered by the role of human-induced processes<br />

linked to global change, especially to l<strong>and</strong> use changes as a driving force <strong>and</strong> primary sector<br />

activities as key factors. This study was conducted at oak (Quercus pubescens Mill.) <strong>and</strong> holm<br />

oak (Quercus ilex L.) mixed forest patches located at Montiferru mountains (Sardinia, Italy).<br />

The aim was to analyze l<strong>and</strong> use & l<strong>and</strong> cover changes during the last five decades, especially<br />

reforestation due to forestl<strong>and</strong> use decrease <strong>and</strong> ab<strong>and</strong>onment circumstances linked to<br />

traditional rural economies’ crisis. <strong>Change</strong>s observed have become apparent at different<br />

biodiversity levels. Firstly, at ecological level mixed forestl<strong>and</strong> surface has grown by mean 13%<br />

<strong>and</strong> reforestation reached upper levels (ca. 25-30%) in areas characterized by low use recurrence.<br />

Secondly, at species level despite the increase of deciduous oaks population in contrast to<br />

previous decades, evergreen oaks tends to dominate mixed forest composition <strong>and</strong> regeneration<br />

dynamics.<br />

Keywords: L<strong>and</strong> use & cover changes, reforestation, mixed forest, Sardinia<br />

1. Introduction<br />

1.1 The role of l<strong>and</strong> use changes into l<strong>and</strong>scape dynamics<br />

Classically l<strong>and</strong> cover dynamics had been exclusively considered under biologist criteria but<br />

nowadays this order is altered by human-induced processes in terms of global change (Turner et<br />

al. 1995), especially by l<strong>and</strong> use <strong>and</strong> its changes as key factors due to contemporaneous human<br />

transformation capacity oversize (Vitousek et al. 1997; Folke et al. 2004). Recent Mediterranean<br />

mountain l<strong>and</strong>scape dynamics are highly determined by two basic premises related to l<strong>and</strong> use<br />

& cover changes. Firstly, the evolution of socioeconomic driving forces merge into strong l<strong>and</strong><br />

use changes during the last half of XX th century, the most important is the decrease, <strong>and</strong> in most<br />

cases, ab<strong>and</strong>onment of primary sector activities. Secondly, the main response of these<br />

l<strong>and</strong>scapes is the increase of forestl<strong>and</strong> surface. Some authors (Rudel et al. 2005) explain this<br />

pattern as the last part of the forest transition; in that paradigm forestl<strong>and</strong> surface becomes<br />

historically reduced until a no-return point when trend reverses <strong>and</strong> forestl<strong>and</strong> cover begins to<br />

increases due to remove l<strong>and</strong> use from disturbance regime (Perry 1998; Folke et al. 2004). One<br />

of the principal phenomena derived from forest transition is reforestation defined as the<br />

reestablishment of forest cover either naturally (by natural seedling, coppice, or root suckers) or<br />

artificially (by direct seedling or planting) (IUFRO Silva Term database). In that sense, natural<br />

reforestation is the most habitual l<strong>and</strong> cover change related to l<strong>and</strong> use change in some<br />

Mediterranean mountain socioecosystems like the Montiferru mountains at Sardinia (Italy)<br />

(Longitude 8º 33’ a 8º 39’ E; Latitude 40º 08’ a 40º 11’ N) where this study was conducted.<br />

* Corresponding author. Tel.: +34935812532 - Fax: +34935813331<br />

Email address: franciscojavier.gomez@uab.es<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F.J. Gómez et al. 2010. L<strong>and</strong> use changes <strong>and</strong> mixed forest dynamics. The case of Montiferru Mountains<br />

371<br />

1.2 Brief Montiferru’s environmental history<br />

The Montiferru Mountains are an archetypical Mediterranean mountain socioecosystem due to<br />

its biodiversity <strong>and</strong> environmental history. Its contemporaneous development starts at XIX th<br />

century when early signs of non domestic l<strong>and</strong> uses increase become apparent, mainly due to<br />

navy <strong>and</strong> pre-industrial activities (glass, leather workshops, etc.) (Beccu 2000). During this<br />

period, characterized by organic economy strong dependent of wood for construction <strong>and</strong><br />

firewood <strong>and</strong> charcoal for fuel, forest raw material consumption grows drove by the expansion<br />

of iron <strong>and</strong> steel industries, railway <strong>and</strong> sea transport <strong>and</strong> due to population growth <strong>and</strong> its<br />

dem<strong>and</strong>. Linked to these <strong>and</strong> other factors like the increase of l<strong>and</strong> plough derived from<br />

agricultural <strong>and</strong> subsistence crisis, forestl<strong>and</strong> covers become highly reduced in all Italy <strong>and</strong><br />

especially in mountain systems like Sardinian ones at latest XIX th <strong>and</strong> the beginning of XX th<br />

century.<br />

First signs of l<strong>and</strong> use shift scenario, based on industrial <strong>and</strong> inorganic-based model, arrived in<br />

the decade of 1930-40 when firewood <strong>and</strong> charcoal consumption drops while fossil fuels<br />

increases (Agnoletti 2003), this trend arrives to mountain socioecosystems later meaning the<br />

beginning of rural economies crisis. It was then when changes in primary sector activities<br />

become apparent, showed in forestry in two general ways; models specialized in fast-growth<br />

wood <strong>and</strong> dedicated to paper <strong>and</strong> pulp emergent industry rise <strong>and</strong> the most extended traditional<br />

uses linked to slow-growth wood, firewood <strong>and</strong> charcoal decrease <strong>and</strong> in most cases become<br />

ab<strong>and</strong>oned few decades later. Other indicators, like rural exodus processes, reinforce this<br />

socioeconomic scenario shift characterized by traditional rural economies’ crisis; at Montiferru<br />

population decreased ca. 24% since 1960s (Mura 1995).<br />

In the last decades of XX th century the massif shows incipient signs of another socioeconomic<br />

scenario shift like the increase of biodiversity management <strong>and</strong> protection measures <strong>and</strong> the rise<br />

of third economic activities. As a result, nowadays the Montiferru still conserves aspects of its<br />

traditional Mediterranean mountain area character but post-industrial socioeconomic scenario<br />

tends to be predominant.<br />

2. Methodology<br />

2.1 Object of study<br />

In the massif of Montiferru oak (Quercus pubescens Mill.) <strong>and</strong> holm oak (Quercus ilex L.)<br />

mixed forest patches occurs in the geographic transition from the upper mesomediterranean to<br />

submediterranean wet-temperate climate conditions, in the altitudinal range comprised between<br />

750 <strong>and</strong> 950 meters a.s.l. In the context of Mediterranean basin, these forestl<strong>and</strong> covers begin its<br />

actual dynamics after Youger Dryas (11.000 B.C.) <strong>and</strong> undergo different fluctuations in its<br />

composition until the frontier between Subboreal <strong>and</strong> Subatlantic (aprox. 3.000 B.C.), when<br />

palinological records show an evident decline of Q.pubescens <strong>and</strong> a increase of Q.ilex (Suc<br />

1984). Two hypotheses try to explain this event; biotic arguments related to the capacity of<br />

holm oak to compete with other trees, <strong>and</strong> the consideration of human-induced effects (Pons <strong>and</strong><br />

Quézel 1985; Barbero et al 1990). At this point, centuries of human intervention on<br />

Mediterranean forestl<strong>and</strong> upsets the balance between oak <strong>and</strong> holm oak populations, favoring<br />

Q.ilex mainly due to its fuel value <strong>and</strong> its food value for cattle in detriment of Q.pubescens,<br />

taxon that in the case of Sardinia <strong>and</strong> Corsica becomes very reduced (Gamisans 1977). In that<br />

sense, more or less isolated character of mixed forest areas contributes to assume the role of<br />

these entities as indicators of global change manifestations, especially of l<strong>and</strong> use changes<br />

processes derived from socioenomic scenario shifts (Lomolino, 2000).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F.J. Gómez et al. 2010. L<strong>and</strong> use changes <strong>and</strong> mixed forest dynamics. The case of Montiferru Mountains<br />

372<br />

2.2 Material <strong>and</strong> methods<br />

The analysis of l<strong>and</strong> use <strong>and</strong> l<strong>and</strong> cover change processes requires the ecological <strong>and</strong> social<br />

criteria integration, linking socioeconomical <strong>and</strong> biophysical driving forces, so as to unravel<br />

l<strong>and</strong>scape dynamics complexity; in this aim hybrid methodologies based on holistic approaches<br />

are basic tools to interpret them (Naveh 2000). In our case, there has been posed a diachronic<br />

model <strong>and</strong> a synchronic model to analyze l<strong>and</strong> use & cover changes.<br />

Diachronic analysis comprises historical records related to l<strong>and</strong> use <strong>and</strong> cartography for the<br />

period 1965-2006. Information about forest use was taken from the review of Corpo <strong>Forest</strong>ale<br />

Regionale of Sardinia archive. Historical series was, in general, vague <strong>and</strong> unsystematic; for this<br />

reason data was treated to homogenize records <strong>and</strong> to set (i) year <strong>and</strong> type of intervention<br />

(familiar or commercial thinning) (ii) species <strong>and</strong> quantity of wood or firewood removed (using<br />

dendrometric rates for Sardinian Quercus sp. <strong>and</strong> normalizing non explicit records into habitual<br />

oak <strong>and</strong> holm oak cutting shift documented for the Montiferru). Besides, oral sources of<br />

information were consulted to improve environmental history of the studied sites. Several<br />

interviews were made to social actors related to Montiferru forestl<strong>and</strong> sphere, especially<br />

primary sector workers (woodcutters, shepherds, etc.) <strong>and</strong> forest owners <strong>and</strong> traders; a wide<br />

majority of people interviewed were elderly men, owners of non written <strong>and</strong> usually non enough<br />

considered information, but very valuable, regarding to relations between human <strong>and</strong><br />

environment. Four sampling stations were selected attending to differences on l<strong>and</strong> use patterns<br />

<strong>and</strong> classified in terms of intensity, considered high when commercial thinning is predominant<br />

<strong>and</strong> low when are related to domestic use (up to 5 m³ per year); <strong>and</strong> recurrence, considered high<br />

when there were more than 20 interventions during recorded period (1965-2006) <strong>and</strong> low in the<br />

other cases (Table 1). Besides, diachronic model comprises the analysis of aerial photographs of<br />

1955, 1977 <strong>and</strong> 2006. Spatial characteristics of mixed forestl<strong>and</strong> patches corresponding to<br />

sampling stations <strong>and</strong> its dynamics have been analyzed by GIS during this period.<br />

Synchronic model comprises the analysis about mixed forestl<strong>and</strong> mass’ ecological dynamics.<br />

For this objective, 20 experimental 10x10 meters plots were r<strong>and</strong>omly distributed into each<br />

sampling station. In each plot, number <strong>and</strong> diameter at breast height (dbh) of oak <strong>and</strong> holm oak<br />

(live <strong>and</strong> dead trees) were measured <strong>and</strong> classified into thee size class: saplings (dbh


F.J. Gómez et al. 2010. L<strong>and</strong> use changes <strong>and</strong> mixed forest dynamics. The case of Montiferru Mountains<br />

373<br />

Table 2: <strong>Change</strong>s in mixed forestl<strong>and</strong> area 1955-2006<br />

Sampling<br />

station<br />

<strong>Forest</strong> patches<br />

<strong>Forest</strong> gaps<br />

1955-1977<br />

(hectares)<br />

1977-2006<br />

(hectares)<br />

balance<br />

(ha)<br />

% of change<br />

(1955-2006)<br />

1955-2006<br />

(hectares)<br />

MA 6,70 7,18 13,88 28,2 -2,28 -52,0<br />

BM -19,50 0,18 -19,32 -25,3 4,36 69,8<br />

SP -0,17 12,66 12,49 26,2 -1,91 -30,4<br />

AS -3,92 8,25 4,33 7,1 -1,23 -16,4<br />

% of change<br />

(1955-2006)<br />

At species level, 299 oaks <strong>and</strong> 1.673 holm oaks live trees were counted <strong>and</strong> measured (per 45<br />

<strong>and</strong> 430 death, respectively). Widespread, holm oak presents higher observed tree densities <strong>and</strong><br />

upper values of basal area (BA) in all the sampling stations than oak that are by mean older;<br />

also, another trend in species composition, documented by the oral sources of information, sets<br />

that nowadays oak density is higher than in previous decades. For l<strong>and</strong> use regimes, the largest<br />

BA <strong>and</strong> average diameter of trees (<strong>and</strong> consequently age) are related to high use intensity.<br />

However, certain values of density <strong>and</strong> basal area take remarkable records in low use recurrence<br />

<strong>and</strong> ab<strong>and</strong>oned areas (SP & MA) where reforestation is higher (Table 3).<br />

Table 3: Estimated mean values for Q.pubescens (Qp) <strong>and</strong> Q.ilex (Qi) density, BA, dbh <strong>and</strong> age<br />

specie MA BM SP AS<br />

Density Qp 485 (330) 270 (211) 295 (216) 445 (349)<br />

(ind/ha) Qi 2.140 (850) 1.580 (494) 2.065 (980) 2.580 (886)<br />

BA Qp 7,23 (8,66) 8,78 (8,75) 5,50 (4,52) 8,22 (5,64)<br />

(m²/ha) Qi 22,33 (8,96) 33,74 (12,71) 28,82 (13,55) 26,36 (10,87)<br />

Mean dbh Qp 7,86 (11,34) 15,2 (12,08) 9,88 (11,56) 12,92 (8,56)<br />

(cm) Qi 7,26 (8,98) 12,5 (10,48) 7,68 (10,22) 8,16 (6,88)<br />

Mean age Qp 18,8 28,7 21,1 24,7<br />

(yr) Qi 17,4 24,6 18,3 18,5<br />

In terms of age structure, significant differences become apparent between use recurrence for<br />

both species (Mann Whitney test for Qp, Z=-2,824 p=0,0047; <strong>and</strong> Qi, Z=-2,786, p=0,0053) due<br />

to younger age size class: low use recurrence areas presents higher sapling frequencies than pole<br />

ones in contrast to high use recurrence regimes where this pattern reverses (Figure 1). Also, as<br />

proposed indicator of mature forest structure (Barbour et al. 1987) areas of high use recurrence<br />

have better adjustments to reverse J-shaped curves than low use ones where reforestation is<br />

higher. The Differences observed in tree-mortality related to l<strong>and</strong> use regime were statistically<br />

non-significant in both species (Kruskal-Wallis test for Qp, H=1,378, df=3, p=0,7106; <strong>and</strong> Qi,<br />

H=7,804, df=3, p=0,0502)<br />

Related to regeneration dynamics 139 oak <strong>and</strong> 1.706 holm oak total ramets were counted (12<br />

<strong>and</strong> 115 death ones, respectively), <strong>and</strong> 666 Q.pubescens <strong>and</strong> 1.298 Q.ilex seedlings too.<br />

As a rule, holm oak presents more resprouting capacity than oak (Mann Whitney test, Z=-7,399,<br />

p


F.J. Gómez et al. 2010. L<strong>and</strong> use changes <strong>and</strong> mixed forest dynamics. The case of Montiferru Mountains<br />

374<br />

related to adult trees presence (Mann Whitney test for Qp Z=-0,775, p= 0,4385; <strong>and</strong> for Qi Z=-<br />

1,587, p=0,1124). For l<strong>and</strong> use regimes, differences between areas in terms of seedlings/plot<br />

ratios are significant (Kruskal-Wallis test for Qp, H=4,476, df=3, p


F.J. Gómez et al. 2010. L<strong>and</strong> use changes <strong>and</strong> mixed forest dynamics. The case of Montiferru Mountains<br />

375<br />

increase of deciduous oaks populations regarding decades ago. In terms of global change<br />

(Peñuelas et al 2007) <strong>and</strong> assuming the uncertainty linked to the complexity of the phenomenon,<br />

it could indicate competitive exclusion episodes of deciduous oaks as shift biome processes.<br />

In conclusion, the analysis of driving forces of l<strong>and</strong> use & cover process highlights limiting<br />

factors <strong>and</strong> conservation risks <strong>and</strong> make it clear the need to integrate l<strong>and</strong> use as an active<br />

management tool to maintain Mediterranean mountain l<strong>and</strong>scapes <strong>and</strong> its biodiversity. However,<br />

in actual global change scenario more future researches are necessary to improve our knowledge<br />

about biodiversity responses, especially to l<strong>and</strong> use <strong>and</strong> climate change components, as a way to<br />

guarantee its conservation.<br />

References<br />

Agnoletti, M., 2003. Bosques e industria de la Madera en Italia, de la unificación al fascismo<br />

(1861-1940). In J.A.Sebastián <strong>and</strong> R. Uriarte (Eds): Historia y economía del bosque en la<br />

Europa del Sur (s. XVIII-XX). Zaragoza: Monografías de Historia Rural. Seminario de<br />

Historia Agraria. Prensas Universitarias de Zaragoza.<br />

Barbero, M. Bonin, G. Loisel, R. Quézel, P., 1990. <strong>Change</strong>s <strong>and</strong> disturbances of forest<br />

ecosystems caused by human activities in the western part of the Mediterranean basin.<br />

Vegetatio, 87:151-173.<br />

Barbour, M.G. Burk, J.H. Pitts, W.D.,1987. Terrestrial Plant Ecology. Benjamin/Cummings<br />

Menlo Park, California.<br />

Beccu, E., 2000. Tra cronaca e storia le vicende del patrimonio boschivo della Sardegna. Carlo<br />

Delfino Editore. Roma.<br />

Folke, C. Carpenter, S. Walker, B. Scheffer, M. Elmqvist, T. Gunderson, L. Holling, C.S., 2004.<br />

Regime shifts, resilience, <strong>and</strong> biodiversity in ecosystem management. Annu. Rev. Ecol.<br />

Syst, 34 :557-581.<br />

Gamisans, J., 1977. La végétation des montagnes corses. Quatrième partie. Phytocoenologia,<br />

4(3): 317-376.<br />

IUFRO Silva Term database (on web). http://www.iufro.org/science/special/silvavoc/silvaterm/<br />

Lomolino, M.V., 2000. A call for a new paradigm of isl<strong>and</strong> biogeography. <strong>Global</strong> Ecology <strong>and</strong><br />

Biogeography, 9:1-6.<br />

Mura, B., 1995. La popolazione in Sardegna. Dati e proiezioni dal 1962 al 2000.Vol IV. Num.<br />

7. Osservatorio Economico e Finanziario della Sardegna.Oristano<br />

Naveh, Z., 2000. What is holistic l<strong>and</strong>scape ecology A conceptual introduction. L<strong>and</strong>scape <strong>and</strong><br />

Urban Planning, 50:7-26.<br />

Perry, D.A., 1998. The scientific basis of forestry. Annual Review of Ecology <strong>and</strong> Systematics,<br />

29:435-466.<br />

Peñuelas, J. Ogaya, R. Boada, M. Jump. A.S., 2007. Migration, invasion <strong>and</strong> decline: changes in<br />

recruitment <strong>and</strong> forest structure in a warming-linked shift of European beech forest in<br />

Catalonia (NE Spain). Ecography, 30(6): 829-837.<br />

Pons, A. Quézel, P., 1985. The history of flora <strong>and</strong> vegetation <strong>and</strong> past <strong>and</strong> present human<br />

disturbance in the Mediterranean area. In C.Gómez-Campo (Ed). Plant conservation in<br />

the Mediterranean area. Dordrecht: Junk Pub.<br />

Rudel, T.K. Coomes, O.T. Moran, E. Achard, F. Angelsen, A. Xu, J. Lambin, E., 2005. <strong>Forest</strong><br />

transitions: towards a global underst<strong>and</strong>ing of l<strong>and</strong> use change. <strong>Global</strong> Environmental<br />

<strong>Change</strong>, 15:23-31.<br />

Suc, J.P., 1984, Origin <strong>and</strong> evolution of the Mediterranean vegetation <strong>and</strong> climate in Europe.<br />

Nature 307(5950): 429-432.<br />

Tuner II, B.L. Gómez-Sal, A. González-Bernáldez, F. Di Castri, F. (Eds), 1995. <strong>Global</strong> L<strong>and</strong><br />

Use <strong>Change</strong>. A perspective from the Columbia Encounter. CSIC. Madrid.<br />

Vitousek, P.M. Money, H.A. Lubchenco, J. Melillo, J.M., 1997. Human domination of Earth’s<br />

ecosystems. Science, 277: 494-499.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L. Leite & J.S.V. Filho2010. Analysis of rainfall in the Vila Velha State Park, State of Parana, Southern Brazil<br />

376<br />

Analysis of rainfall in the Vila Velha State Park, State of Parana,<br />

Southern Brazil, in the period between 1954 <strong>and</strong> 2001<br />

Maysa de Lima Leite 1* & Jorim Sousa das Virgens Filho 2<br />

1 Departamento de Biologia Geral, Universidade Estadual de Ponta Grossa, Av. Carlos<br />

Cavalcanti, 4748, 84.030-900, Paraná, Brazil<br />

2 Departamento de Matemática e Estatística, Universidade Estadual de Ponta Grossa,<br />

Paraná, Brazil<br />

Abstract<br />

The aim of this study was to perform a temporal analysis involving frequency, intensity <strong>and</strong><br />

variability of rainfall for the city of Ponta Grossa, between 1954 <strong>and</strong> 2001. Data of daily rainfall<br />

(mm) were analyzed <strong>and</strong> obtained from the Meteorological Station, situated in the area of the<br />

State Park of Vila Velha, Paraná State, Southern Brazil. The Park constitutes a Conservation<br />

Unit with a total area of 3122.11 hectares <strong>and</strong> presents vegetation of open grassl<strong>and</strong> <strong>and</strong><br />

scattered clumps of forest, with the focus on the Paraná Pine (Araucaria angustifolia). The<br />

annual average rainfall observed for the series was 1546.2 mm, revealing a growing trend over<br />

the years. The month with the highest average rainfall was January <strong>and</strong> the lowest average was<br />

observed in August. Summer was the most rainy season, including the highest number of days<br />

with rain. The range of “drizzle” showed the highest frequency in all months.<br />

Keywords: temporal analysis, intensity, variability, rainfall.<br />

1. Introduction<br />

Rainfall is one of the meteorological elements that most influence on environmental conditions<br />

<strong>and</strong> has a great importance because it is directly related to many sectors of the society, since the<br />

rainfall affects the economy, the environment <strong>and</strong> the society as a whole (Silva et al., 2007).<br />

The spatial-temporal distribution of rainfall is a very important regional characteristic, both for<br />

society <strong>and</strong> for the economy. Knowledge of this feature can guide decisions on the measures<br />

necessary to minimize the damage from irregular rainfall (Piccinini, 1993 apud Minuzzi et al.,<br />

2007). According to Nery et al. (2002 apud Nery et al., 2004), information about the number of<br />

days with rainfall are useful both in agricultural planning in a short-term (which agronomic<br />

practices for soil <strong>and</strong>/or air moisture are conditioning factors), as a long-term conditions<br />

(definitions of regions <strong>and</strong> times most suitable for the sowing of crops <strong>and</strong>/or for maintenance<br />

of perennial species). To Assis (1991 apud Ribeiro <strong>and</strong> Lunardi, 1997), the frequency, i.e. the<br />

number of days within a month or season in which rainfall occurs, perhaps is the most important<br />

aspect in relation to rain for vegetation in general, besides quantity <strong>and</strong> variability. The<br />

objective of this study was to evaluate the frequency <strong>and</strong> intensity of rainfall from the<br />

Meteorological Station located in the Vila Velha State Park, near the city of Ponta Grossa, State<br />

of Parana, Southern Brazil, over the period of 1954 to 2001.<br />

* Corresponding author. Tel.: +55 (42) 3220 3126 Fax: +55 (42) 3220-3102<br />

Email address: mleite@uepg.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L. Leite & J.S.V. Filho2010. Analysis of rainfall in the Vila Velha State Park, State of Parana, Southern Brazil<br />

377<br />

2. Methodology<br />

Daily rainfall data (mm) were obtained from the Meteorological Station, belonging to the<br />

Instituto Agronômico do Paraná (IAPAR), located in the State Park of Vila Velha, on the<br />

geographical coordinates of 25°13'S, 50°01' W <strong>and</strong> 880 m of altitude. The data were collected<br />

from January 1954 to December 2001, constituting a series of 48 years. The State Park of Vila<br />

Velha, has an area of 3,122.11 hectares <strong>and</strong> is situated in the city of Ponta Grossa, State of<br />

Paraná, Southern Brazil (see Figure 1). The main access is the BR-376, an important road<br />

junction that connects the shores, Curitiba (the State capital) <strong>and</strong> the North of the Paraná State.<br />

Ponta Grossa is located in the Second Parana Plateau <strong>and</strong> is situated in an area of grassl<strong>and</strong><br />

(short grass steppe) with clumps of forest of Araucaria. Because of this phytogeographical<br />

feature, the region where Ponta Grossa is located is called the Campos Gerais Region of Paraná<br />

State (Maack, 1981). Initially, the daily rainfall data were subjected to screening <strong>and</strong> assessment<br />

of consistency of time series. To assess whether there was any particular trend related to the<br />

annual totals of rainfall, there was used the concept of moving average, calculated by cycles of<br />

five years. After that, it was proceeded the frequency analisys of the days with rain in each<br />

month of the series. The days with climatological drought (Leaves & Fisch, 2006) were<br />

excluded <strong>and</strong> the remaining data were divided into class intervals of rainfall. From there, they<br />

were classified in relation to the intensity of cumulative daily rainfall as: drizzle (0.1 to 2.5 mm),<br />

light rain (2.5 to 10.0 mm), moderate rain (10.0 to 25.0 mm), heavy rain (from 25.0 to 50.0 mm)<br />

<strong>and</strong> extreme rain (above 50 mm) (Calvetti et. al., 2006). It was also evaluated the annual totals<br />

of rainfall <strong>and</strong> annual totals of days with rain <strong>and</strong> climatological drought.<br />

Figure 1: Location of Vila Velha State Park<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L. Leite & J.S.V. Filho2010. Analysis of rainfall in the Vila Velha State Park, State of Parana, Southern Brazil<br />

378<br />

3. Result<br />

Figure 2 shows the annual totals of rainfall (mm) for the city of Ponta Grossa (PR) from 1954<br />

to 2001. It is observed that the year with the higher total of rainfall was 1998 with 2494 mm,<br />

<strong>and</strong> the year with the lowest total was 1985, with 910.3 mm. For the series studied, it was<br />

obtained an annual average total of rainfall of 1546.2 mm. The estimate of the moving average<br />

for the total annual rainfall, considering cycles of five years throughout the series, evidenced the<br />

existence of a positive trend over the annual totals of rainfall, as explained by the regression line<br />

with the time variable (r 2 = 0.5264 ). The graph in Figure 1 also shows the alternation between<br />

high <strong>and</strong> low annual totals of rainfall, which can partly be explained by the occurence of El<br />

Niño <strong>and</strong> La Niña. Analyzing the data in a monthly scale, it can be observed that the month with<br />

the highest average rainfall was January (185.4 mm) <strong>and</strong> the month with the lowest average was<br />

August (78.9 mm). Considering the occurrence of rain, the year 1983 was the year with more<br />

rainy days (168 days), while 1968 was the year that had the lowest total of days with rain (only<br />

73 days). The average for the studied period was 126 days with rain per year, showing also a<br />

tendency to increase this number of days in the course of time.<br />

Rainfall (mm)<br />

2700<br />

2500<br />

2300<br />

2100<br />

1900<br />

1700<br />

1500<br />

1300<br />

1100<br />

900<br />

1635<br />

1638<br />

1504<br />

1432<br />

2080 1302 1308<br />

1731<br />

1578<br />

968<br />

925<br />

Annual totals of rainfall (mm)<br />

r² = 0,5264<br />

1537<br />

1420<br />

1637<br />

1701<br />

1539<br />

1515<br />

1381 2217 1362<br />

2055<br />

1532<br />

1567<br />

1605<br />

1330<br />

1301<br />

1180<br />

910<br />

1927<br />

2494<br />

1839<br />

1415<br />

2000<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

Rainfall (mm) - Moving Average<br />

Total rainfall (mm) Moving average - 5 years Linear (Moving average - 5 years)<br />

Figure 2. Annual totals <strong>and</strong> moving averages of rainfall (mm) over the period of 1954 to 2001.<br />

On Table 1, which classifies the rainfall intensity, it can be seen the percentage that each class<br />

interval represents, over the total number of days with rainfall, for each month of the year. The<br />

predominant class interval throughout the year is light rain (2.5 - 10.0 mm accumulated in a<br />

day), being the most frequent during nine months of the year, ranging from 29.46% attendance<br />

in June to 33.66 % in March. The range of drizzle (0.1 - 2.5 mm) was more frequent in April<br />

(30.08%) <strong>and</strong> the range of moderate rainfall (10 - 25 mm) was the most frequent in October<br />

(29.76%). In May, drizzle <strong>and</strong> light rain had the same percentage (29.11%). The range of heavy<br />

rain (25 - 50 mm) ranged from 8.73% of frequency in November to 14.79% in April. With<br />

respect to the rain considered extreme (over 50 mm a day), it was observed that this range was<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L. Leite & J.S.V. Filho2010. Analysis of rainfall in the Vila Velha State Park, State of Parana, Southern Brazil<br />

379<br />

less frequent in all months, exceeding 5% frequency only in May (5.57%) <strong>and</strong> with the lowest<br />

frequency in February (1.18%).<br />

Table 1. Percentage corresponding to the class intervals of rainfall over the total number of days with<br />

rainfall for each month analyzed. CRI = Classification of Rainfall Intensity; RI = Rainfall Intensity.<br />

CRI RI (mm) J (%) F (%) M (%) A (%) M (%) J (%)<br />

Drizzle 0,1 ├ 2,5 28,47 28,11 31,23 30,08 29,11 26,73<br />

Light rain 2,5 ├ 10,0 31,11 32,25 33,66 27,57 29,11 29,46<br />

Moderate rain 10,0 ├ 25,0 25,42 26,18 23,14 25,81 24,05 25,50<br />

Heavy rain 25,0 ├ 50,0 11,94 12,28 9,55 14,79 12,15 13,37<br />

Extreme rain > 50,0 3,06 1,18 2,43 1,75 5,57 4,95<br />

TOTAL 100,00 100,00 100,00 100,00 100,00 100,00<br />

CRI RI (mm) J (%) A (%) S (%) O (%) N (%) D (%)<br />

Drizzle 0,1 ├ 2,5 25,78 27,11 22,45 24,86 26,79 28,34<br />

Light rain 2,5 ├ 10,0 33,14 32,23 33,06 28,49 32,14 31,76<br />

Moderate rain 10,0 ├ 25,0 24,65 26,51 26,94 29,76 30,16 26,87<br />

Heavy rain 25,0 ├ 50,0 12,46 12,35 14,29 13,97 8,73 9,77<br />

Extreme rain > 50,0 3,97 1,81 3,27 2,90 2,18 3,26<br />

TOTAL 100,00 100,00 100,00 100,00 100,00 100,00<br />

4. Discussion<br />

The phenomenon called El Niño Southern Oscillation (ENSO), considered as the main cause of<br />

climate variability in various regions of the globe is a phenomenon of ocean-atmosphere<br />

interaction that occurs in the tropical Pacific Ocean <strong>and</strong> has two extreme phases: a warm phase<br />

known as El Niño <strong>and</strong> a cold phase called La Niña (Berlato & Fontana, 2003). Among the<br />

consequences of El Niño, is the increase in rainfall in southern South America, reaching<br />

catastrophic proportions, as occurred in 1983 <strong>and</strong> 1998. In those years, there were registered the<br />

two largest annual totals of rainfall in the series studied in Ponta Grossa (2494 mm in 1998 <strong>and</strong><br />

2217 mm in 1983). The other three largest totals of rainfall (2080.1 mm in 1957, 2071.1 mm in<br />

1993 <strong>and</strong> 2054.7 mm in 1990) also occurred in El Niño years, although that the event was not as<br />

strong as in 1983 <strong>and</strong> 1998. Furthermore, the La Niña phenomenon, opposite to El Niño, which<br />

corresponds to the anomalous cooling of surface waters in the equatorial Pacific, Central <strong>and</strong><br />

Eastern, may explain the lowest rainfall observed in the series studied (910.3 mm in 1985).<br />

From the st<strong>and</strong>point of hydrology, rainfall up to 10 mm have little impact, except to moisten the<br />

soil to the point of reaching its field capacity, increasing the efficiency of runoff into rivers.<br />

Despite of the reduced frequency observed, the “very heavy” rains can cause further damage<br />

such as l<strong>and</strong>slides, floods, etc.. as a result of the use <strong>and</strong> occupation of l<strong>and</strong>. It is very important<br />

to make a suitable plan for l<strong>and</strong> use, paying particular attention to areas of risk, as close to river<br />

banks or slopes, which are greatly affected sites when it rains heavily. The uncertainty related to<br />

global changes connected to precipitation, originated from natural phenomenas <strong>and</strong>/or<br />

anthropogenic activities, makes it currently impossible to establish categorically the effects of<br />

climate change on ecosystems <strong>and</strong> on agricultural activities more broadly. Nevertheless, studies<br />

on various scales, including regional scale, will enable the underst<strong>and</strong>ing of how natural<br />

ecosystems can respond <strong>and</strong> adapt to these climatic variability, becoming an increasingly<br />

urgent need. Identified potential vulnerabilities, one should start the search for strategies <strong>and</strong><br />

technologies for adaptation, including taking advantage of possible climate changes that may be<br />

considered beneficial.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L. Leite & J.S.V. Filho2010. Analysis of rainfall in the Vila Velha State Park, State of Parana, Southern Brazil<br />

380<br />

Acknowlegments<br />

The authors thank to the Instituto Agronômico do Paraná (IAPAR) for the authorization to<br />

access the rainfall data, <strong>and</strong> to the SETI – Fundação Araucaria, by the research support.<br />

References<br />

Berlato, M.A. <strong>and</strong> Fontana, D.C., 2003, El Niño e La Niña: impactos no clima, na vegetação e<br />

na agricultura do Rio Gr<strong>and</strong>e do Sul; aplicações de previsões climáticas na agricultura.<br />

Porto Alegre: Editora da UFRGS. 110 p.<br />

Calvetti, L., Beneti, C., Gonçalves, J.E., Moreira, I.A., Duquia, C., BREDA, A., ALVES, T.A.,<br />

2006. Definição de classes de precipitação para utilização em previsões por categoria e<br />

hidrológica. In: Congresso Brasileiro de Meteorologia, 14, Florianópolis, Anais...<br />

Florianópolis, SBMet.:100-105.<br />

Folhes, M.T., Fisch, G., 2006. Caracterização climática e estudo de tendências nas séries<br />

temporais de temperatura do ar e precipitação em Taubaté (SP). Revista Ambiente e Água –<br />

An interdisciplinary journal of Applied Science, 1, n. 1: 61-71.<br />

Maack, R. Geografia física do Estado do Paraná, 1981. 2. ed. Rio de Janeiro: J. Olympio;<br />

Curitiba: Secretaria da Cultura e do Esporte do Governo do Estado do Paraná, 450 p.<br />

Minuzzi, R.B., Sediyama, G.C., Barbosa, E.M., Melo Júnior, J.C.F., 2007. Climatologia do<br />

comportamento do período chuvoso da Região Sudeste do Brasil. Revista Brasileira de<br />

Meteorologia, Rio de Janeiro, 22, n. 3: 338-344.<br />

Nery, J.T., Martins, M.L.O.F., Roseghini, W.F.F., Roseghini, F.F., 2004. Variabilidade da<br />

precipitação pluvial e disponibilidade hídrica na Região Noroeste do Estado do Paraná.<br />

Revista Brasileira de Agrometeorologia, Santa Maria, 12, n. 2: 289-297.<br />

Ribeiro, A.M.A., Lunardi, D.M.C.,1997. A precipitação mensal provável para Londrina – PR,<br />

através da função Gama. Energia na Agricultura, Botucatu, 12, n. 4: 37-44.<br />

Silva, J.C., Heldwein, A.B., Martins, F.B., Trentin, G., Grimm, E.L., 2007. Análise de<br />

distribuição de chuva para Santa Maria, RS. Revista Brasileira de Engenharia Agrícola e<br />

Ambiental, Campina Gr<strong>and</strong>e, 11, n. 1: 67–72.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L.Leite & J.S.V. Filho 2010. Identification of climatic trends for some localities in the Southern Region of Campos Gerais<br />

381<br />

Identification of climatic trends for some localities in the Southern<br />

Region of Campos Gerais <strong>and</strong> surroundings, State of Parana, Brazil,<br />

through the analysis of historical data of rainfall <strong>and</strong> temperature<br />

Maysa de Lima Leite 1* & Jorim Sousa das Virgens Filho 2<br />

1 Departamento de Biologia Geral, Universidade Estadual de Ponta Grossa, Av. Carlos<br />

Cavalcanti, 4748, 84.030-900, Paraná, Brazil<br />

2 Departamento de Matemática e Estatística, Universidade Estadual de Ponta Grossa,<br />

Paraná, Brazil<br />

Abstract<br />

This study aimed to assess possible changes in climate localities in the southern region of<br />

Campos Gerais (Fern<strong>and</strong>es Pinheiro, Lapa, Ponta Grossa), Parana, Brazil. The forest vegetation<br />

occupies 22% of the area of Campos Gerais, including different types <strong>and</strong> successional stages.<br />

These forests are naturally fragmented, forming isolated groves of various sizes <strong>and</strong> extensions,<br />

located on the slopes, small depressions or tracks that come in rivers, streams <strong>and</strong> springs. To<br />

investigate possible changes in climate, daily temperature <strong>and</strong> precipitation data were analyzed,<br />

by the use of Mann-Kendall test. Data analysis revealed an increase in average <strong>and</strong> minimum<br />

temperatures in the three localities. For the maximum temperature, the city of Ponta Grossa<br />

showed negative trend, a fact explained by the increase in cloudiness in the region. For the<br />

rainfall data, the city of Ponta Grossa <strong>and</strong> Fern<strong>and</strong>es Pinheiro showed a positive trend, while for<br />

the city of Lapa, it was negative.<br />

Keywords :precipitation, temperature, climatic trends<br />

1. Introduction<br />

The climate is characterized by the interaction between various climatic variables, especially<br />

among precipitation <strong>and</strong> temperature. Both are closely related <strong>and</strong> these changes may cause<br />

major changes in regional climate. There are few studies on long-term variability of extreme<br />

weather <strong>and</strong> climate in Brazil <strong>and</strong> South America. In addition, studies in some regions of Brazil,<br />

or in the rest of South America, have used different methodologies, which does not allows<br />

intercomparisons or geographical integration. The lack of meteorological information of good<br />

quality related to daily series in large areas of Brazil, <strong>and</strong> the very restricted access to daily<br />

weather information stored in databases of meteorological services, have not allowed<br />

identification of climatic extremes <strong>and</strong> their variability, especially in tropical South America. To<br />

the South of Brazil <strong>and</strong> Northern Argentina, the works of Marengo <strong>and</strong> Camargo (2006) <strong>and</strong><br />

Rusticucci <strong>and</strong> Barruc<strong>and</strong> (2004) showed negative trends in diurnal temperature range due to<br />

positive trends registered in minimum temperature. They also observed an increased frequency<br />

of hot days in winter. Compared to the air temperature, a larger number of studies of trends in<br />

precipitation have been developed due to the greater availability of data on rainfall than<br />

temperature. Groisman et al. (2005) found positive trends of systematic increases of rainfall <strong>and</strong><br />

extremes of rainfall in the subtropical region, in the South <strong>and</strong> Northeast of Brazil. The authors<br />

considered that the Southeast, since 1940, has shown systematic increases in the frequency of<br />

heavy rainfall, up almost 58% in recent years. This study aimed to examine the climatological<br />

* Corresponding author. Tel.: +55 (42) 3220 3126 - Fax: +55 (42) 3220 3102<br />

Email address: mleite@uepg.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L.Leite & J.S.V. Filho 2010. Identification of climatic trends for some localities in the Southern Region of Campos Gerais<br />

382<br />

variables, temperature <strong>and</strong> precipitation <strong>and</strong> check for possible changes <strong>and</strong> climate trends<br />

observed in the cities of Fern<strong>and</strong>es Pinheiro, Lapa <strong>and</strong> Ponta Grossa, Southern of the Campos<br />

Gerais Region <strong>and</strong> surroundings, State of Parana, Brazil.<br />

2. Methodology<br />

The cities of Ponta Grossa <strong>and</strong> Lapa are located in the phytogeographical region known as<br />

Campos Gerais of Paraná, while Fern<strong>and</strong>es Pinheiro is located in a surrounding area, Paraná<br />

State, Southern Brazil. These cities are represented on the map in Figure 1.<br />

Figure 1 – Location of the cities: Fern<strong>and</strong>es Pinheiro, Lapa <strong>and</strong> Ponta Grossa, State of Parana, Southern<br />

Brazil.<br />

The region of Campos Gerais holds unique l<strong>and</strong>scapes such as escarpments, caves, canyons,<br />

rivers with rocky beds, waterfalls, ruiniform reliefs, remarkable displays of rocks <strong>and</strong> fossils<br />

<strong>and</strong> diverse floral species. The forest vegetation occupies about 22% of the area of Campos<br />

Gerais, including different types <strong>and</strong> successional stages. Such forests have been naturally<br />

fragmented, forming isolated copses of various sizes <strong>and</strong> extensions, located on the slopes,<br />

small depressions or in b<strong>and</strong>s accompanying rivers, creeks <strong>and</strong> springs. However, the<br />

occupation of areas previously considered unproductive, the rise of mechanized agriculture <strong>and</strong><br />

the use areas for the production of wood, have become factors that could cause preoccupation,<br />

<strong>and</strong> cause, over time, major changes in the regional ecosystem, requiring studies of diverse<br />

natures. The data used in this study were provided by the Instituto Agronômico do Paraná<br />

(IAPAR) <strong>and</strong> consisted of records of the climatological variables temperature (maximum,<br />

minimum <strong>and</strong> average) (ºC) <strong>and</strong> precipitation (mm). Initially, the daily data were subjected to<br />

screening <strong>and</strong> assessment of consistency of time series <strong>and</strong> then arranged in annual, quarterly<br />

<strong>and</strong> monthly series. The nonparametric test of Mann-Kendall, initially proposed by Sneyers<br />

(1975), was applied in a significance level of 0.05 <strong>and</strong> 0.01%, in order to identify possible<br />

climate changes over time, which could be presented as positive or negative trends within the<br />

period analyzed. The data series have different durations for each city <strong>and</strong> location details can<br />

be found in Table 1.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L.Leite & J.S.V. Filho 2010. Identification of climatic trends for some localities in the Southern Region of Campos Gerais<br />

383<br />

Table 1 . Length of series <strong>and</strong> location of the municipalities of Fern<strong>and</strong>es Pinheiro, Lapa <strong>and</strong> Ponta<br />

Grossa – State of Parana, Brazil.<br />

City Latitude Longitude Altitude (m) Series<br />

Fern<strong>and</strong>es<br />

Pinheiro<br />

Number of<br />

years<br />

-25°27’ -50°35’ 893 1963-2007 45<br />

Lapa -25°47’ -49°46’ 910 1989-2007 19<br />

Ponta Grossa -25°13’ -50°01’ 880 1954-2001 48<br />

3. Result<br />

The climatological data were analyzed with the nonparametric Mann-Kendall test, producing<br />

the results shown in Table 2.<br />

Table 2 . Results of the Mann-Kendall Test applied on the climatological data of rainfall <strong>and</strong> temperature<br />

in the cities of Fern<strong>and</strong>es Pinheiro, Lapa <strong>and</strong> Ponta Grossa, State of Parana, Brazil.<br />

Localidade Variável Teste de Mann-<br />

Kendall (U calculado)<br />

0,05% de<br />

significância<br />

0,01% de<br />

significância<br />

Fern<strong>and</strong>es Precip itação 0,35 NS NS<br />

Pinheiro<br />

Fern<strong>and</strong>es Temperatura<br />

3,30 S S<br />

Pinheiro Máxima<br />

Fern<strong>and</strong>es Temperatura<br />

5,81 S S<br />

Pinheiro Média<br />

Fern<strong>and</strong>es Temperatura<br />

5,38 S S<br />

Pinheiro Mínima<br />

Lapa Precip itação -1,92 NS NS<br />

Lapa<br />

Lapa<br />

Lapa<br />

Ponta<br />

Grossa<br />

Ponta<br />

Grossa<br />

Ponta<br />

Grossa<br />

Ponta<br />

Grossa<br />

Temperatura<br />

Máxima<br />

2,41 S NS<br />

Temperatura<br />

2,27 S NS<br />

Média<br />

Temperatura<br />

2,27 S NS<br />

Mínima<br />

Precip itação 2,06 S NS<br />

Temperatura<br />

Máxima<br />

Temperatura<br />

Média<br />

Temperatura<br />

Mínima<br />

-3,18 S S<br />

1,88 NS NS<br />

4,85 S S<br />

With respect to the rainfall data, a positive trend was observed for the city of Ponta Grossa,<br />

while for the other two cities, the results were not estatistically significant. For temperatures<br />

(maximum, average <strong>and</strong> minimum), significant trends were detected for all the cities, at least<br />

with 0.05% of significance. For the maximum temperature there was a positive trend for the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L.Leite & J.S.V. Filho 2010. Identification of climatic trends for some localities in the Southern Region of Campos Gerais<br />

384<br />

cities of Lapa <strong>and</strong> Fern<strong>and</strong>es Pinheiro, the latter being significant at 0.01%, while for the city of<br />

Ponta Grossa, the maximum temperature had a significant negative trend of 0.01%. For<br />

minimum temperature, all the cities had significant positive trends, showing a rise in the<br />

minimum temperature in the region. These results of a temperature rise in the southern region of<br />

Campos Gerais, in the State of Parana, is reinforced by analyzing the average temperature in the<br />

three localities, which showed positive trends. Finally, for the city of Ponta Grossa, in despite of<br />

having observed an average temperature rise, this was not significant for the studied period.<br />

4. Discussion<br />

According to Silva (2004) the rainfall in southern Brazil are distributed regularly throughout the<br />

year, compared to other regions of the country, but in recent years the phenomenon El Nino has<br />

served continuously in this region. This climatic phenomenon, when actuates, significantly<br />

alters the levels of rainfall, which are increased during the year. Whereas this phenomenon has<br />

been acting steadily in the region, the increase in the totals of annual rainfall <strong>and</strong> the positive<br />

trend observed for the city of Ponta Grossa may be correlated with the performance of such a<br />

frequent phenomenon in the region. Back (2001) also found elevated totals of annual rainfall for<br />

the city of Urussanga, State of Santa Catarina, identifying a significant positive trend. The<br />

decrease of the maximum temperature in Ponta Grossa, along with higher minimum temperature,<br />

can be partially explained as a consequence of increased cloudiness in the region. Due to this<br />

rise in the levels of cloudiness during the day, a smaller amount of sunlight can reach the earth's<br />

surface, occurring a decrease in maximum temperature. During the night, with lots of clouds,<br />

there is greater retention of radiation, complicating its loss to space, thereby increasing the<br />

minimum temperature at location (Silva & Guetter, 2003). In a previous study, Silveira & Gam<br />

(2006), analyzing changes in minimum temperatures in southern Brazil, found that the<br />

minimum temperature in the state of Rio Gr<strong>and</strong>e do Sul also showed a positive trend over the<br />

study period under review. The same increases were verified by Obregon & Marengo (2007),<br />

which developed a major study on climate throughout the Brazilian territory, in order to<br />

determine possible changes in climate variables. In this study the annual minimum temperatures<br />

in the city of Curitiba, Brazil, showed a significant positive trend at 0.05%, reinforcing the<br />

findings of elevated temperatures in the State of Parana. One reason for this increase in<br />

temperatures can also be correlated to the increase in the totals of areas destinated to<br />

urbanization in the last years (Sansigolo et al., 1992). In some regions, deforestation <strong>and</strong><br />

changes in l<strong>and</strong> use, as a result of human activities have increased rapidly in recent decades, <strong>and</strong><br />

there are evidences that these actions can modify the thermodynamic characteristics of the<br />

lower atmosphere. These changes are the result of complex interactions between climate,<br />

hydrology, vegetation <strong>and</strong> management of water resources <strong>and</strong> l<strong>and</strong>.<br />

Acknowlegments<br />

The authors thank to the Instituto Agronômico do Paraná (IAPAR) for the authorization to<br />

access the rainfall data, <strong>and</strong> to the SETI – Fundação Araucaria, by the research support.<br />

References<br />

Back, J.A., 2001. Aplicação de análise estatística para a identificação de tendências climáticas.<br />

Pesq. Agropec. Bras., 36, n. 5: 717-726.<br />

Groisman, P., Knight, R., Easterling, D., Karl, T., Hegerl, G., Razuvaev, V., 2005. Trends in<br />

tense precipitation in the climate record. Journal of Climate, 18: 1326-1350.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L.Leite & J.S.V. Filho 2010. Identification of climatic trends for some localities in the Southern Region of Campos Gerais<br />

385<br />

Marengo, J., Alves, L., Camargo, H., 2005. An overview of global climate predictability at<br />

seasonal to interannual time scales. <strong>Global</strong> Energy <strong>and</strong> Water Cycle Experiment -<br />

GEWEX Newsletter, 15, n. 4: 5-7.<br />

Obregon, G. & Marengo, J.A., 2007. Caracterização do clima no século XX no Brasil:<br />

Tendências de chuvas e temperaturas médias extremas. CPTEC/INPE, 4: 1-3.<br />

Rusticucci, M., Barruc<strong>and</strong>, M., 2004: Observed trends <strong>and</strong> changes in temperature extremes<br />

over Argentina. Journal of Climate, 17: 4099-4107.<br />

Sansigolo, C., Rodriguez, R., Chichury, P., 1992. Tendências nas temperaturas médias do Brasil.<br />

In: Congresso Brasileiro de Meteorologia, 7, São Paulo. Anais. São Paulo: 367-371.<br />

Silva, M.E.S. & Guetter, A.K., 2003. Mudanças Climáticas regionais observadas no estado do<br />

Paraná. Terra Livre., 1, n. 20: 111-126.<br />

Silva, I.R., 2001. Variabilidade sazonal e interanual das precipitações na região sul do Brasil<br />

associadas às temperaturas dos oceanos Atlântico e Pacífico. São José dos Campos,<br />

SP: INPE, 98p.<br />

Silveira, V.P. & Gam, M.A., 2006. Estudo de tendências das temperaturas mínimas na Região<br />

Sul do Brasil. In: Congresso Brasileiro de Meteorologia, 14, Florianópolis. Anais.<br />

Florianópolis: 189-195.<br />

Sneyers, R. Sur l.analyse statistique des series d.observations, 1975. Genève : Organisation<br />

Météorologique Mondial, (OMM Note Technique, 143), 192 p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Ortega et al. 2010. Monitoring vulnerability of the Spanish forest l<strong>and</strong>scapes: the SISPARES approach<br />

386<br />

Monitoring vulnerability of the Spanish forest l<strong>and</strong>scapes: the<br />

SISPARES approach<br />

Marta Ortega 1* , Valentín Gomez 1 , Jose Manuel García del Barrio 2 & Ramon Elena-<br />

Rosselló 1<br />

1 University School of <strong>Forest</strong> Technical Engineers, Polytechnic University of Madrid,<br />

Spain<br />

2 National Institute of Research <strong>and</strong> Agrarian <strong>and</strong> Food Technology, Madrid, Spain<br />

Abstract<br />

SISPARES is a system devoted to study the ecological evolution of the Spanish Rural<br />

<strong>L<strong>and</strong>scapes</strong> including their characterisation, classification, <strong>and</strong> recent past <strong>and</strong> future changes. It<br />

is based on a sampling network of 215 permanent 4x4 km plots surveyed by using aerial<br />

photographs of three dates: 1956, 1984 <strong>and</strong> 1998. Currently, a new survey is going on. The<br />

network design was based on an environmental stratification of Spain well linked to the<br />

European Environmental Stratification.<br />

<strong>Forest</strong> L<strong>and</strong>scape Vulnerability (FLV) is defined as the risk of human disturbance, especially<br />

forest fires. We assessed FLV using an index based on <strong>Forest</strong> L<strong>and</strong>scape Fragility (FLF)<br />

weighted by Road Accessibility. FLF is defined as the risk of l<strong>and</strong>scape disturbance caused by a<br />

high degree of interspersion <strong>and</strong> juxtaposition between patches of forest <strong>and</strong> non forest l<strong>and</strong><br />

uses: The non forest l<strong>and</strong> uses, such urban, crop <strong>and</strong> managed grassl<strong>and</strong>, are considered as<br />

potential sources of human disturbances, for the forest l<strong>and</strong> use patches, including woodl<strong>and</strong>,<br />

“matorral”, “dehesa”, tree plantation <strong>and</strong> riparian woodl<strong>and</strong>.<br />

From 1956 till 1998, a significant increase of FLV has been observed in SISPARES l<strong>and</strong>scape<br />

plots. This increase is related to the increment of forest fires in Spain during the last 40 years.<br />

Keywords: L<strong>and</strong>scape, monitoring, vulnerability, SISPARES system, Spain<br />

1. Introduction<br />

L<strong>and</strong>scape vulnerability is defined as the risk of severe disturbances <strong>and</strong> could be consider as a<br />

character of l<strong>and</strong>scape patterns (composition <strong>and</strong> configuration). In recent years, l<strong>and</strong>scape<br />

character assessment in Europe has become central to sustainable development <strong>and</strong> the<br />

management of l<strong>and</strong>. It is recognised as an important tool for policy stakeholders, which<br />

provides them with quantitative <strong>and</strong> qualitative evidence to reach a dynamic management,<br />

adjustable to new dem<strong>and</strong>s of regional identity (Washer 2005). But only, Denmark (Nordic<br />

Council of Ministers 1987), Austria (Wrbka et al. 1999) <strong>and</strong> Spain (Elena-Rosselló et al. 2005)<br />

have developed national approaches for delimitation <strong>and</strong> mapping local spatial units on the basis<br />

of a range of l<strong>and</strong>scape features (geophysical, cultural <strong>and</strong> historic, perception <strong>and</strong> aesthetics<br />

<strong>and</strong> natural value) as a basis for vulnerability assessment. However, in many other countries all<br />

over the world, local approaches exist <strong>and</strong> some of them have been specifically designed to<br />

evaluate l<strong>and</strong>scape vulnerability to forest fire (e.g. Preston et al. 2009, Sorrensen 2009). They<br />

are important tools for preventing the risk of forest fire.<br />

* Corresponding author.<br />

Email address: marta.ortega.quero@upm.es<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Ortega et al. 2010. Monitoring vulnerability of the Spanish forest l<strong>and</strong>scapes: the SISPARES approach<br />

387<br />

In Spain, the SISPARES national approach is an ongoing project designed to study the<br />

ecological value <strong>and</strong> dynamics of rural l<strong>and</strong>scapes in Spain, including their characterisation <strong>and</strong><br />

classification (Elena-Rosselló et al. 2005). <strong>Forest</strong> L<strong>and</strong>scape Vulnerability (FLV) is among of<br />

the ecological l<strong>and</strong>scape characteristics assessed <strong>and</strong> then monitored. FLV has been defined as<br />

the “risk of human disturbance” <strong>and</strong> measured as “the interspersion among forested <strong>and</strong> non<br />

forested patches weighted by road accessibility”. According to its definition, we might<br />

hypothesize that FLV is a good indicator of forest fire risk. Therefore, it is necessary to prove<br />

that hypothesis by studying its relationship with forest fire incidence.<br />

This paper compares the FLV values in Spain among 1956 to 1998 with the annual national<br />

burnt area <strong>and</strong> forest fire frequency (Spanish Environmental Ministry 2005) trying to find out<br />

their relationship type. An increase of FLV could be determining a higher number of forest fires<br />

<strong>and</strong>/or an increase of burnt areas.<br />

2. Material <strong>and</strong> methods<br />

The vulnerability of Spanish forest l<strong>and</strong>scapes has been monitored by using the SISPARES<br />

approach which has a stratified simple sampling design based on the Biogeoclimatic L<strong>and</strong><br />

Classification of Spain, known by its acronym CLATERES (Elena-Rossello et al. 1997). This<br />

stratification was constructed using a divisive multivariate classification approach adapted from<br />

the Country Survey l<strong>and</strong> classification system (Bunce et al. 1996a, 1996b, 1996c), applied to<br />

climatic, physiographic <strong>and</strong> geological data. The initial stage was the establishment of a<br />

representative Spanish Rural L<strong>and</strong>scape Network (REDPARES) that has 206 4x4 km squares in<br />

Iberian Peninsula <strong>and</strong> Balearic Isl<strong>and</strong>s. All the squares were surveyed using aerial photographs<br />

at three dates: 1956, 1984 <strong>and</strong> 1998 to derive measurements of 11 major l<strong>and</strong> cover (Table 1). A<br />

new survey is currently in progress to updating the samples by interpretation of air photos from<br />

2007, but the data are not let ready to be presented in this work.<br />

Each square of SISPARES represents the l<strong>and</strong>scape of one geoclimatic class <strong>and</strong> was analysed<br />

by delimiting patches of l<strong>and</strong> cover <strong>and</strong> linear elements of road network from aerial photo<br />

interpretation <strong>and</strong> field surveys during the nineties. The scale of photos is 1:30.000 <strong>and</strong> the<br />

minimum patch size that it has been interpreted is 1 ha. The patches are relatively homogeneous<br />

portions of l<strong>and</strong> that represent different l<strong>and</strong> covers that are adjacent <strong>and</strong> make up the l<strong>and</strong>scape.<br />

Table 1: Types of l<strong>and</strong> cover detected by interpretation of aerial photographs <strong>and</strong> their correspondence<br />

with l<strong>and</strong> cover CORINE classification (EEA 1995)<br />

Type of l<strong>and</strong> cover<br />

CORINE/level<br />

<strong>Forest</strong><br />

Matorral<br />

Dehesa<br />

<strong>Forest</strong> plantation<br />

Pastures<br />

Crops<br />

Riparian woodl<strong>and</strong><br />

Rock<br />

Water body<br />

Urban <strong>and</strong> industrial use<br />

<strong>Forest</strong>/3.1<br />

Shrub <strong>and</strong>/or herbaceous vegetation<br />

associations/3.2<br />

Agro-forestry areas/2.4.4<br />

Young Broad-leave forests/3.1.1. <strong>and</strong> Young<br />

Coniferous forests/3.1.2<br />

Pastures/2.3.1<br />

Arable l<strong>and</strong>/2.1 <strong>and</strong> permanent crops/2.2<br />

Riparian woodl<strong>and</strong>/3.1.1.5<br />

Bare rock/3.3.2<br />

Water bodies/5.1.2<br />

Artificial surfaces/1<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Ortega et al. 2010. Monitoring vulnerability of the Spanish forest l<strong>and</strong>scapes: the SISPARES approach<br />

388<br />

2.1 Data analysis<br />

In each square FLV has been calculated as:<br />

FLV=FLF*RD/100 (1)<br />

Where RD is road density in the square <strong>and</strong> FLF is an index of forest l<strong>and</strong>scape fragility, which<br />

has been calculated as:<br />

FLF=IJI*WFN (2)<br />

Where IJI is an index of interspersion <strong>and</strong> juxtaposition that considers the neighbourhood<br />

relations between patches. Each patch is analysed for adjacency with all other patch types <strong>and</strong><br />

measures the extent to which patch types are interspersed i.e. equally bordering other patch<br />

types. The IJI is a relative index that represents the observed level of interspersion as a<br />

percentage of the maximum possible given the total number of patch types (McGarigal et al.<br />

1995).<br />

m = number of patch types,<br />

(3)<br />

E ik = length of edge between patch type i <strong>and</strong> patch type k.<br />

IJI approaches 0 when the distribution of adjacencies among unique patch types becomes<br />

increasingly uneven. IJI = 100 when all patch types are equally adjacent to all other patch types<br />

(i.e., maximum interspersion <strong>and</strong> juxtaposition.<br />

WFN in equation 2 is the contrast between forest <strong>and</strong> non forest areas in the square <strong>and</strong> is<br />

measured as:<br />

(1-| <strong>Forest</strong> area – Non <strong>Forest</strong> area|)/100 (4)<br />

A l<strong>and</strong>scape square will be more fragile when a higher WFN <strong>and</strong> a higher IJI have.<br />

Besides if the square has a higher road density it will be a high value of vulnerability.<br />

Fragstat software was used to calculate the IJI index. Repeated measures ANOVA was used to<br />

analyse FLV, FLF, IJI, WFN <strong>and</strong> RD of 206 squares of SISPARES in three dates: 1956, 1984<br />

<strong>and</strong> 1998 by means of Statistica softhware.<br />

3. Results <strong>and</strong> discussion<br />

FLV index showed a significant increase in Spain along the study period (F=3.36; p=0.04) as it<br />

is observed in the figure 1. The increase between 1956 <strong>and</strong> 1984 was 2% <strong>and</strong> during the last 24<br />

years of this study period the mean number of forest fires was 3,940 per year (Table 2).<br />

Between 1984 <strong>and</strong> 1998 data, the increase of FLV was 12% more <strong>and</strong> only during these 14<br />

years the mean number of forest fires was 15,844 per year. In the same way the burnt areas were<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Ortega et al. 2010. Monitoring vulnerability of the Spanish forest l<strong>and</strong>scapes: the SISPARES approach<br />

389<br />

near twice over in the second period, mainly caused by the increase in the mean of burnt non<br />

forest areas per year (Table 2).<br />

Per zones, FLV has been higher in northwest of Spain because to a typical complex l<strong>and</strong>scape<br />

structure that provides higher levels of l<strong>and</strong>scape fragility, interspersion of patch types <strong>and</strong><br />

contrast between forest <strong>and</strong> non forest areas (Fig 1a). Lower FLV has been observed in aridity<br />

Figure 1: Monitoring of <strong>Forest</strong> L<strong>and</strong>scape Vulnerability (FLV) in 206 squares of SISPARES approach<br />

along three dates: (a) 1956, (b) 1984 <strong>and</strong> (c) 1998. The point size indicates the value of FLV.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Ortega et al. 2010. Monitoring vulnerability of the Spanish forest l<strong>and</strong>scapes: the SISPARES approach<br />

390<br />

zones because to a typical agricultural l<strong>and</strong>scape that determines low fragility, interspersion of<br />

patch types <strong>and</strong> contrasts.<br />

But this framework of 1956 changed with the ab<strong>and</strong>on of crops productions <strong>and</strong> the increase of<br />

forest plantation between 1956 <strong>and</strong> 1984 mainly (Ortega et al., 2008). These changes of l<strong>and</strong><br />

use has produced a decrease of FLF along the study period (F=17.55; p


M. Ortega et al. 2010. Monitoring vulnerability of the Spanish forest l<strong>and</strong>scapes: the SISPARES approach<br />

391<br />

over time <strong>Forest</strong> Systems (Investigación Agraria: Sistemas y Recursos <strong>Forest</strong>ales), 17(2),<br />

114-129<br />

Preston, B.L., Brooke, C., Measham, T. G., Smith, T. F. <strong>and</strong> Gorddard, R, 2009. Igniting<br />

change in local government: lessons learned from a bushfire vulnerability assessment.<br />

Mitig Adapt Strateg Glob <strong>Change</strong>, 14: 251–283.<br />

Sorrensen, C., 2009. Potential hazards of l<strong>and</strong> policy: Conservation, rural development <strong>and</strong> fire<br />

use in the Brazilian Amazon. L<strong>and</strong> Use Policy, 26: 782-791.<br />

Spanish Environmental Ministry, 2005. Los incendios forestales de España. Centro de<br />

coordinación de la información nacional sobre incendios forestales. 106 pp.<br />

Wascher, D.M. (ed). 2005. European L<strong>and</strong>scape Character Areas – Typologies, Cartography<br />

<strong>and</strong> Indicators for the Assessment of Sustainable <strong>L<strong>and</strong>scapes</strong>. Wageningen, The<br />

Netherl<strong>and</strong>s. Alterra Report No.1254, 150 pp.<br />

Wrbka, T., Szerencsits, E., Reiter, K.<strong>and</strong> Plutzar, C. 1999. Which attributes of l<strong>and</strong>scape<br />

structure can be used as indicators for sustainable l<strong>and</strong> use Experiences from alpine <strong>and</strong><br />

lowl<strong>and</strong> l<strong>and</strong>scapes in Austria. In: Kover, P. et al. (Eds) Nature <strong>and</strong> culture in l<strong>and</strong>scape<br />

ecology. Proceedings of the CZ-IALEConference “Present <strong>and</strong> historical nature-culture<br />

interactions in l<strong>and</strong>scapes – experiences for the 3 rd Millennium”. Prague, Charles<br />

University: 80–94.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Puddu et al. 2010. L<strong>and</strong>scape transformations seen through the historical cartography: Sardinia as case study<br />

392<br />

L<strong>and</strong>scape transformations seen through the historical cartography:<br />

Sardinia as case study<br />

Giuseppe Puddu 1,2* , Raffaele Pelorosso 2 , Federica Gobattoni 2 & Maria Nicolina Ripa 2<br />

1<br />

Monterano Regional Reserve, Canale Monterano, Italy<br />

2 Department of Environmental <strong>and</strong> <strong>Forest</strong>ry (D.A.F.), Tuscia University, Viterbo, Italy<br />

Abstract<br />

L<strong>and</strong> Cover <strong>Change</strong>s are directly linked to l<strong>and</strong>scape transformations due to human activities. In<br />

the last century l<strong>and</strong> changes dynamics have increased just like the attention towards the need of<br />

l<strong>and</strong>scape conservation. L<strong>and</strong>scape, as a collector of “environmental objects <strong>and</strong> relations<br />

existing between them”, can be considered a privileged point of view in order to underst<strong>and</strong> the<br />

territorial dynamics. <strong>Forest</strong>ry systems, as biodiversity collectors, have suffered the biggest<br />

changes in terms of loss/increase of surface. This work regarding Sardinian isl<strong>and</strong>, can be<br />

considered the first one, at this scale, about changes analysis focused on forestry l<strong>and</strong>scape.<br />

Recovering historical cartography, <strong>and</strong> integrating the obtained data with different data-sources,<br />

a new frame is designed to underline the increase of forestry surface. The obtained results could<br />

be a reliable <strong>and</strong> useful tool to direct new studies on forestry l<strong>and</strong>scape, especially for<br />

Mediterranean <strong>and</strong> Apennine regions <strong>and</strong> to manage new scenarios on biodiversity conservation.<br />

Keywords: Sardinia, L<strong>and</strong> Use & L<strong>and</strong> Cover <strong>Change</strong>, <strong>Forest</strong>ry l<strong>and</strong>scape, Historical<br />

maps, Deforestation,<br />

1. Introduction<br />

Sardinia is the second largest isl<strong>and</strong> in the Mediterranean (after Sicily) with a total area of<br />

roughly 24,000 km 2 . With a resident population of more than 1,600,000 inhabitants (43% of<br />

which concentrated in two main urban areas), Sardinia is one of the Italian regions with the<br />

lowest demographic density. The isl<strong>and</strong> has a complex topography, with more than 80% of the<br />

region occupied by hilly <strong>and</strong> mountainous areas (>300 m a.s.l.), <strong>and</strong> with a maximum elevation<br />

of 1,834 m a.s.l. The highest human population densities occur in the main plains where large<br />

agricultural areas are also located. In Sardinia, forests had always played a key role with an<br />

essential importance for local community to strengthen the isl<strong>and</strong> culture <strong>and</strong> identity <strong>and</strong> to<br />

reflect <strong>and</strong> spread the “idea” of Sardinia as a flourishing l<strong>and</strong> in people coming from foreign<br />

l<strong>and</strong>s. <strong>Forest</strong>s have been seen as a refuge <strong>and</strong>, on the opposite side, as also a precious natural<br />

resource, as “wood”, leading to a conflict between the different interests in forest floor<br />

exploitation since the first travels through the Mediterranean Sea by navigators such as the<br />

Phoenician people. The presence of woodl<strong>and</strong> has always been considered as an obstacle to<br />

extensive <strong>and</strong> subsistence farming but also to the intensive one that ran over plains <strong>and</strong> hills on<br />

the isl<strong>and</strong> arousing the need to reduce <strong>and</strong> destroy forests where l<strong>and</strong>s could be managed, with a<br />

more or less income, for agriculture. In the meantime, forest mantles were the appropriate space<br />

to support the presence of herds during the dry summer periods that isl<strong>and</strong> climate usually<br />

shows as, <strong>and</strong> often, lead to the cut of lower branches to get green forage until rare <strong>and</strong><br />

* Corresponding author. Tel.: +39 0761 357359 - Fax: +39 0761 357250<br />

Email address: puddu.foresta@tiscali.it<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Puddu et al. 2010. L<strong>and</strong>scape transformations seen through the historical cartography: Sardinia as case study<br />

393<br />

umbrella-shaped formations took place <strong>and</strong> became a typical character of Sardinia inl<strong>and</strong> as<br />

they are many times cited to represent the old residue of primordial formations or ancient<br />

Mediterranean forests in the Isl<strong>and</strong>. According to Le Lannou (1941) “There are few regions, in<br />

Europe, where natural green l<strong>and</strong>scapes- if with this term we refer to l<strong>and</strong>scapes that had not<br />

changed under l<strong>and</strong> cultivation- have a so important role as in Sardinia.” All different observers<br />

writing on Sardinia, describe the nineteenth century as a century that deeply altered the<br />

“complex l<strong>and</strong>scape” in the Isl<strong>and</strong>, changing it from a l<strong>and</strong> covered by woods with a mosaic of<br />

surfaces dedicated to sheep farming, passing through management models characterized by a<br />

dichotomy between agriculture <strong>and</strong> forestry, to much more sophisticated agro-silvo-pastoral<br />

systems. Even if it’s not clearly delineated in l<strong>and</strong>scaping, this issue has been quite focused in<br />

terms of l<strong>and</strong>scape, if this word refers to a “wide container of objects <strong>and</strong> systemic<br />

relationships”, since woods <strong>and</strong> forest have always been complex territorial objects, with a<br />

multitasking character covering private <strong>and</strong> “public heritage” concerns even if under a private<br />

property. L<strong>and</strong> governing coming from legislation (with dichotomies right/wrong in the<br />

question statement <strong>and</strong> in the imposed solutions, <strong>and</strong> applied/not-applied for what regarding the<br />

proper realization of laws expectations on territory), is the actual source of observed reality,<br />

giving evidence of social movements <strong>and</strong> economical drivers replying to legal acts.<br />

2. Methodology<br />

To underst<strong>and</strong> the actual l<strong>and</strong>scape it’s then necessary to ensure a diachronic interpretation of it<br />

through time because only with the comprehension of territorial motivations of the past, the<br />

dynamics <strong>and</strong> the identity of the present can be really realized <strong>and</strong> acquired. This historical<br />

lecture of the presence <strong>and</strong> distribution in forests on the isl<strong>and</strong>, turns to fix a new attempt of<br />

analysis in comparison to the what have been done <strong>and</strong> is archived in literature, <strong>and</strong> sets the aim<br />

of this work. In Sardinia, L<strong>and</strong>scape Issues (Conservation <strong>and</strong> Management) had often assumed<br />

the distinguishing feature of “Identity Definition” in community talks <strong>and</strong> social living during<br />

the end of nineteenth century <strong>and</strong> he beginning of the twentieth, leading to a contrast between<br />

two different economic scenarios for forest management:<br />

1. <strong>Forest</strong> cut was configured as a generalized deforestation, with an undue removing of<br />

woods <strong>and</strong> a disfigurement of Sardinia<br />

2. <strong>Forest</strong> cut was configured a san economical operation needed to increase the<br />

government income <strong>and</strong> to let modernity come in the Isl<strong>and</strong> updating its services to the<br />

rest of the isl<strong>and</strong><br />

2.1.1 The “<strong>Forest</strong> Map” made by National <strong>Forest</strong> Militia<br />

To analyze the effects of forest management at regional scale along a time period of more than a<br />

century (1850-2010), it’s necessary to integrate <strong>and</strong> complete different sources of data so that to<br />

delineate a trend of areas covered by woods as fully as possible. In particular, historical maps,<br />

never analyzed before, have been recovered; these maps have been realized in the period 1930-<br />

35 <strong>and</strong>, called as “<strong>Forest</strong> Map”, reports location <strong>and</strong> composition of woodl<strong>and</strong>s at 1:100000<br />

scale. This cartography is one the first forest inventory, elaborated at a national scale (where<br />

Italy is represented in 267 sheets <strong>and</strong> Sardinia in 26 sheets with an extension of 30 degrees in<br />

longitude <strong>and</strong> 20 degrees in latitude). To restore all the thematic information through<br />

digitalization, single sheets have been georeferenced by “rubber sheeting” with three layers<br />

(coats of the isl<strong>and</strong>, administrative boundaries <strong>and</strong> roads) so that to obtain a great number of<br />

control points (at least 100 points per sheets) <strong>and</strong> produce a direct georeferentiation in WGS84<br />

(Geri et al. 2008). This method, even if it’s not so correct from a formal point of view, allows a<br />

large computational time saving in comparison to the georeferentiation in the native system (not<br />

so clear because of geographical metadata lack) <strong>and</strong> re-projection in other systems <strong>and</strong>, in<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Puddu et al. 2010. L<strong>and</strong>scape transformations seen through the historical cartography: Sardinia as case study<br />

394<br />

particular, (due to the great number of points) has introduced errors that have been estimated<br />

lower than map scale precision.<br />

2.1.2 Map of l<strong>and</strong> utilization made by National Research Counsil, Italian Cadastre, <strong>and</strong><br />

Italian Touring Club.<br />

The second map used in this work, is called “Map of l<strong>and</strong> utilization” realized by a cooperation<br />

between National Research Council (CNR), Italian Cadastre (Catasto) <strong>and</strong> the Italian Touring<br />

Club (TCI) <strong>and</strong> it has been released in 1952-1960. This map can be assimilated to a l<strong>and</strong> use <strong>and</strong><br />

l<strong>and</strong> cover map, even if the legend, made by 21 items, cannot be completely compared with that<br />

one of the European CORINE Project (Falcucci et al. 2007). Different opinions exist about its<br />

original setting <strong>and</strong>, actually, the real thematic aggregation criteria, that led to the final map<br />

(1:200000 scale) starting from all he information contained in the single sheets of the Italian<br />

Cadastre (1:1000; 1:2000; 1: 4000 scales), are not known<br />

2.1.3 The CORINE L<strong>and</strong> Cover map<br />

The third map used for comparisons, is derived from CORINE Project, that, for Italy, produced<br />

a digital map at the nominal scale of 1:100000 with 44 legend items.<br />

2.2 Model development<br />

To develop a model for a diachronic comparison between different cartographic maps, it’s<br />

necessary to obtain the best compatibility between them, even heterogeneous in legend <strong>and</strong><br />

elaboration methods (topographic survey vs aerial <strong>and</strong> satellite photos analysis) that can be<br />

realized according to the procedures of map generalization (Weibel <strong>and</strong> Jones 1998)<br />

summarized in three consequential phases (Petit <strong>and</strong> Lambin 2002; Pelorosso et al. 2009). Due<br />

to the different structure of legends in the maps used, thematic generalization could be obtained<br />

only reducing all the items into two classes: wood vs non-woods. Spatial resolution<br />

generalization was derived by a rasterization on the same DTM at a 20x20 m of resolution. For<br />

every map, consequential aggregations were achieved starting from 40 m cells with a step of 20<br />

m until 500 m cells were reached using the majority rule <strong>and</strong> giving rise to 24 maps for each of<br />

the national cartographies. Using five l<strong>and</strong>scape indexes (Riitters 1995) every aggregated map<br />

can be analyzed <strong>and</strong> a distance can be elaborated in 5-dimensional space of l<strong>and</strong>scape metrics<br />

between the generalized <strong>and</strong> the target map. In this way, the best resolution can be identified to<br />

make a comparison between target map (CLC, 1990) <strong>and</strong> generalized maps (<strong>Forest</strong> Map 1930-<br />

1935; Map of L<strong>and</strong> Utilization 1952-1960). See Table 1:<br />

Table 1. Comparison framework between maps.<br />

<strong>Forest</strong> Map (1930-35) vs Map of L<strong>and</strong> Utilization (1952-60) MIL 040 ÷ MIL 500 → CNR-TCI 040<br />

<strong>Forest</strong> Map (1930-35) vs CORINE (1990) MIL 040 ÷ MIL 500 → CORINE 1990 040<br />

Map of L<strong>and</strong> Utilization (1952-60) vs CORINE (1990) CNR 040 ÷ CNR 500 → CORINE 1990 040<br />

3. Results<br />

The first obtained results, show a great reduction of forest extensions after the Second World<br />

War, in a quite justifiable way. This collapse could be coherent with the extinction risk found<br />

for different animal species linked to forestal ecosystems (e.g. Corsican Red Deer, Fallow Deer,<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Puddu et al. 2010. L<strong>and</strong>scape transformations seen through the historical cartography: Sardinia as case study<br />

395<br />

Golden Eagle). Table 2 shows the results from the analysis, giving evidence to the consistent<br />

decrease occurred between ‘30ies <strong>and</strong> ‘60ies, according to Falcucci et al. 2007 for Sardinia.<br />

Table 2. First results from map comparisons.<br />

Cartographic data <strong>Forest</strong> extensions %<br />

Milizia (1930-35) 2,473.9 km 2<br />

CNR-TCI (1952-60) 1,911.9 km 2 - 22,7<br />

Corine L<strong>and</strong> Cover IT (1990) 3,853.5 km 2 101,6<br />

Corine L<strong>and</strong> Cover IT (2000) 3,879.2 km 2 0,7<br />

Corine L<strong>and</strong> Cover Sar (2008) 5,034 km 2 29,8<br />

Figure 1. <strong>Forest</strong> contraction/expansion trend in Sardinia during the twentieth century.<br />

Analyzing maps, two by two, with crosstab operation, <strong>Forest</strong> Map 1930-35 had much more<br />

similarities with CLC-1990 map than with CNR-TCI-Catasto map 1952-1960, in spite of the<br />

less elapsed time (for crosstabulation results, see Figure 2).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Puddu et al. 2010. L<strong>and</strong>scape transformations seen through the historical cartography: Sardinia as case study<br />

396<br />

Figure 2. Results of crosstabulation between map<br />

Recurring to other sources of information analysis, with a particular reference to National<br />

statistics <strong>and</strong> forest studies about Italian southern regions (D’Autilia et al. 1967; Merendi 1954;<br />

Quattrocchi 1950; IFN 1985) an alternative scenario can be delineated, in which the extensions<br />

have not decreased (Table 3), but, on the contrary, they show a slow <strong>and</strong> constant increase as<br />

Figure 3 shows.<br />

Table 2. Other results from comparison between historical maps <strong>and</strong> inventory <strong>and</strong> statistical data.<br />

Cartographic data <strong>Forest</strong> extensions %<br />

<strong>Forest</strong> Map (1930-35) 2,473.9 km 2<br />

Merendi (1954) 2,956.4 19,5<br />

1° National Inventory <strong>Forest</strong> 2,925 km 2 -1,1<br />

Corine L<strong>and</strong> Cover IT (1990) 3,853.5 km 2 101,6<br />

Corine L<strong>and</strong> Cover IT (2000) 3,879.2 km 2 0,7<br />

Corine L<strong>and</strong> Cover Sar (2008) 5,034 km 2 29,8<br />

4. Discussion<br />

Figure 3.<strong>Forest</strong> expansion trend in Sardinia during the twentieth century.<br />

If these interpretation analysis is correct, forests in Sardinia remained constant or increased<br />

along twentieth century, because of the main law on <strong>Forest</strong>s (Law 3267, 1923) that imposed an<br />

accurate direct management of forests. According to what is reported in Beccu, 2000, Sardinia,<br />

before 1877 (when a permissive law on woods-cut was delivered) had more than 4,000 km 2 of<br />

woodl<strong>and</strong> that actually seem to be fully restored.<br />

Table 3. Comparison between forests datasets in Sardinia<br />

Beccu, 2000 (woodl<strong>and</strong> at <strong>Forest</strong> Map, 1930-35 CLC Sardinia, 2008<br />

‘800)<br />

<strong>Forest</strong> 4,2566 km 2 <strong>Forest</strong> 2,473.9 km 2 <strong>Forest</strong> 5,034 km 2<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


G. Puddu et al. 2010. L<strong>and</strong>scape transformations seen through the historical cartography: Sardinia as case study<br />

397<br />

The most known picture of Sardinia in terms of l<strong>and</strong>scape, is linked to pastoral customs <strong>and</strong><br />

habits, in which the aesthetic component is the diffusion of pasture l<strong>and</strong>, <strong>and</strong> the increase of<br />

forests implies a strong environmental change that have to be taken into account in territorial<br />

planning strategies, since it’ll lead to the new (“old”) forest issue of relation between pasture<br />

<strong>and</strong> wood. Given that, in less than a century, the biological <strong>and</strong> ecological drives led forest<br />

systems to occupy spaces <strong>and</strong> territories (when these would be able to support forest growth),<br />

then a remark on “Habitat Directive” model in Sardinia (but also in Mediterranean basin) is due<br />

since a wide protection is reserved to semi-natural open areas, derived from (continuous)<br />

anthropic rehashing, even to the detriment of woods. If the dreaded destruction of forest habitat<br />

seems not to be occurred in 60’s, as several Authors identified as the main cause of extinction<br />

<strong>and</strong>/or reducing in animal biodiversity on the Isl<strong>and</strong>, then the reasons of this extinction should<br />

be better analyzed <strong>and</strong> understood (e.g. trespassing, agriculture industrialization, urban<br />

expansion, industrial development).<br />

References<br />

Beccu, E., 2000. Tra cronaca e storia le vicende del patrimonio boschivo della Sardegna. Carlo<br />

Delfino Editore, Sassari.<br />

D’Autilia, M., Sommazzi, S. <strong>and</strong> Arrigoni, P.V., 1967. Estensione e produzione dei boschi della<br />

Sardegna. Atti Convegno Prosettive economico-industriali della produzione legnosa in<br />

Sardegna, Cagliari. 33-75 p.<br />

Falcucci, A., Maiorano L., <strong>and</strong> Boitani, L., 2007. <strong>Change</strong>s in l<strong>and</strong>-use/l<strong>and</strong>-cover patterns in<br />

Italy <strong>and</strong> their implications for biodiversity conservation. L<strong>and</strong>scape Ecology, 22:617-631.<br />

Geri, F., Giordano, M., Nucci, A., Rocchino, D. <strong>and</strong> Chiarucci A., 2008. Analisi Multitemporale<br />

della Provincia di Siena mediante l’utilizzo di cartografie storiche. <strong>Forest</strong>@, 5:82-91.<br />

Le Lannou, M., 1941. Pâtres et paysans de la Sardaigne. Arrault <strong>and</strong> Cie, Tours, VIII-364 p.<br />

Merendi, A., 1954. Aspetti del problema forestale e montano nel mezzogiorno d’Italia. Italia<br />

<strong>Forest</strong>ale e Montana, IX:129-139.<br />

Pelorosso, R., Leone A. <strong>and</strong> Boccia L., 2009. L<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> use change in the Italian<br />

central Apennines: a comparison of assessment methods. Applied Geography, 29:35-48.<br />

Petit, C.C. <strong>and</strong> Lambin, E.F., 2001. Impact of data integration technique on historical l<strong>and</strong>use/l<strong>and</strong>-cover<br />

change: Comparing historical maps with remote sensing data in the<br />

Belgian Ardennes. L<strong>and</strong>scape Ecology, 17:117-132.<br />

Quattrocchi, G., 1950. Le superfici boscate in Italia al 30 giugno 1947. Istituto Poligrafico dello<br />

Stato, Roma.<br />

Riitters, K.H., O’Neil, R.V., Hunsaker, C.T., Wickham, J.D., Yankee, D.H., Timmins, S.P.,<br />

Jones, K.B. <strong>and</strong> Jackson, B.L., 1995. A factor analysis of l<strong>and</strong>scape pattern <strong>and</strong> structure<br />

metrics. L<strong>and</strong>scape Ecology, 10:23-39.<br />

Weibel, R. <strong>and</strong> Jones, C.B., 1998 Computational perspective on map generalization.<br />

GeoInformatica, 2(4):307-314.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H.R. Ratsimba et al. 2010. Multi-scale analysis of carbon stocks <strong>and</strong> deforestation monitoring<br />

398<br />

Multi-scale analysis of carbon stocks <strong>and</strong> deforestation monitoring –<br />

Case of the Eastern tropical humid forest of Madagascar<br />

Harifidy Rakoto Ratsimba 1 *, L.G. Rajoelison 1 , F.M. Rabenilalana 1 , J. Bogaert 2 &<br />

E. Haubruge 3<br />

1 Département des Eaux et Forêts, Ecole Supérieure des Sciences Agronomiques,<br />

Université d’Antananarivo, Madagascar<br />

2 Service d’Ecologie du Paysage et systèmes de production végétale - Université Libre<br />

de Bruxelles, Belgium<br />

3 Gembloux Agro-Bio Tech, Université de Liège, Belgium<br />

Abstract<br />

Carbon accounting <strong>and</strong> deforestation monitoring constitute the most important<br />

instruments in implementing the Reducing Emission from Deforestation <strong>and</strong> Degradation<br />

(REDD) mechanism in developing countries like Madagascar. Furthermore, most of the<br />

researches <strong>and</strong> calculations have been established at different scales. The present research<br />

develops a methodology to assess the aboveground biomass at a large scale through vegetation<br />

indices. Thus, biomass inventory in the field provides data for calibration <strong>and</strong> classification of<br />

the Normalized Difference Vegetation Index (NDVI) combined with the Enhanced Vegetation<br />

Index (EVI) on Moderate Resolution Imaging Spectroradiometer Image. The results<br />

demonstrate that NDVI <strong>and</strong> EVI provide an accurate classification of forest carbon at a large<br />

scale. Moreover, the use of vegetation indices enables the replication of the classification over<br />

time, useful for the detection of changes <strong>and</strong> the establishment of a deforestation baseline.<br />

Keywords: vegetation indices, tropical humid forest, deforestation monitoring, carbon<br />

accounting, Madagascar.<br />

1. Introduction<br />

The estimation of this last decade shows that tropical forests represent 25% of the terrestrial<br />

biosphere’s carbon (Bonan, 2008). This forest ecosystem plays a very important role in the<br />

global carbon cycle by storing about 80% of all above-ground terrestrial organic carbon (IPCC,<br />

2001). For this reason, the United Nations Framework Convention on Climate <strong>Change</strong> (1998)<br />

<strong>and</strong> its Kyoto Protocol (2003) recognized the role of forests in carbon sequestration. In this way,<br />

the deforestation <strong>and</strong> degradation processes in tropical regions become more <strong>and</strong> more<br />

important: l<strong>and</strong> conversion is the main reason for 93.4% of the annual net forest loss (FAO,<br />

2001). The Bali Action Plan, adopted by UNFCCC at the thirteenth session of its Conference of<br />

the Parties (COP 13 – 2007), m<strong>and</strong>ates Parties to negotiate a post 2012 tool, including possible<br />

incentives for forest-based climate change mitigation actions. This COP 13 has adopted a<br />

decision on “Reducing emissions from deforestation <strong>and</strong> degradation in developing countries” <strong>and</strong><br />

encourages Parties to explore a range of actions, identify options <strong>and</strong> undertake efforts to address<br />

the drivers of deforestation <strong>and</strong> forest degradation. Moreover, it points out the need of<br />

methodologies related to REDD emissions reporting.<br />

Satellite remote sensing technologies are currently widely tested <strong>and</strong> suggested as a tool in<br />

REDD monitoring, reporting <strong>and</strong> verification. However, there is a debate as to the overall<br />

feasibility <strong>and</strong> cost-benefit ratio of remote sensing approaches, depending on the wide range of<br />

*Corresponding author. Tel.: +261 34 08 00 627<br />

Email address: rrharifidy@moov.mg<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H.R. Ratsimba et al. 2010. Multi-scale analysis of carbon stocks <strong>and</strong> deforestation monitoring<br />

399<br />

ecosystem <strong>and</strong> l<strong>and</strong> use conditions as well as the range of approaches to carbon credit<br />

accounting (Holmgren et al., 2008).<br />

This study is carried out in the tropical humid forest of Madagascar which is located in the<br />

eastern part of the isl<strong>and</strong>. The objective is to propose a methodology of upscaling process in<br />

deforestation <strong>and</strong> degradation assessment through satellite images (different resolutions)<br />

analysis <strong>and</strong> biomass inventory in the field. A spatial modelling approach is applied to search<br />

specific factors correlated to biomass stock, to identify vegetation indices in order to simulate<br />

the forest area across large regions, <strong>and</strong> to estimate the deforestation trend.<br />

2. Methodology<br />

2.1. Study area<br />

Madagascar should be considered among the highest conservation priorities (Myers, 1988;<br />

McNeely et al., 1990; Mittermeier et al., 1992). However, the poverty remains one of the main<br />

causes of deforestation <strong>and</strong> degradation. The average income is around $200 per year<br />

(Population Reference Bureau, 1992). Moreover, more than 70 % are from the rural regions<br />

which are facing low agricultural productivity (CIA, 2009). These situations have led to a<br />

continuous deforestation <strong>and</strong> degradation process in the whole country.<br />

The eastern part of Madagascar is characterized by a tropical humid climate with over 2000 mm<br />

mean rainfall per year <strong>and</strong> about 26°C annual mean temperature. The vegetation of this region is<br />

divided into three primary categories corresponding to elevation b<strong>and</strong>s: "lowl<strong>and</strong> rain forest" (0<br />

to 800 m), "moist montane forest" (800 to 1,300 m), <strong>and</strong> "sclerophyllous montane forest" (1,300<br />

to 2,300 m) (White, 1983). The third category is not considered in this study due to its limited<br />

repartition (the patches are to small to be evaluated in a 232 m * 232 m pixel of MODIS).<br />

There are some general trends in forest characteristics with increasing elevation: decreasing<br />

stature, fewer straight unbranched <strong>and</strong> boled trees, less stratification, more epiphytes, more<br />

bryophytes <strong>and</strong> lichens, a better developed <strong>and</strong> more diverse herb layer, <strong>and</strong> floristic changes<br />

(Lewis et al. 1996). Dense evergreen trees characterize the lowl<strong>and</strong> forest up to 800 m, with a<br />

canopy exceeding 30 m. The mid-altitude moist forest is as rich in species as the lowl<strong>and</strong> forest,<br />

but tends to have a shorter canopy of 20 to 25 m.<br />

The main threat is from subsistence agriculture through slash <strong>and</strong> burn activities (“tavy”) <strong>and</strong><br />

illegal logging which are the main causes of deforestation <strong>and</strong> degradation (Humbert, 1927;<br />

Rauh, 1979; Jolly <strong>and</strong> Jolly, 1984; Sussman et al., 1985; Jenkins, 1987).<br />

All of these aspects make REDD monitoring, reporting <strong>and</strong> verification very important for the<br />

country by creating clear <strong>and</strong> simple methods <strong>and</strong> tools to evaluate deforestation threat across a<br />

l<strong>and</strong>scape especially for policy makers <strong>and</strong> forest managers.<br />

2.2. Biomass <strong>and</strong> forest carbon accounting<br />

Biomass is defined as “organic material both above-ground <strong>and</strong> below-ground, <strong>and</strong> both living<br />

<strong>and</strong> dead, e.g., trees, crops, grass, tree litter, roots etc.” (Samalca et al., 2007). The IPCC<br />

(Intergovernmental Panel on Climate <strong>Change</strong>, 2003, 2006), has defined five different carbon<br />

pools for Greenhouse Gas (GHG) inventory : (1) living above-ground biomass (AGB), (2)<br />

living below-ground biomass (BGB), (3) dead organic matter in wood (DOM), (4) dead organic<br />

matter in litter (DOM), <strong>and</strong> (5) soil organic matter (SOM). Biomass is converted into carbon by<br />

multiplying its weight with a carbon fraction of dry matter which is usually 0.5 (IPCC, 2006).<br />

Many biomass estimation researches are focused on above-ground forest biomass (Aboal et al.,<br />

2005; Kraenzel et al., 2003; Laclau, 2003; Losi et al., 2003; Rakoto Ratsimba et al., 2010)<br />

because it represents the majority of the total biomass.<br />

This study focuses on the link between the assessment of above-ground forest biomass based on<br />

Rakoto Ratsimba et al. methods (2010) <strong>and</strong> recorded vegetation indices on satellite images.<br />

2.3. Stratification <strong>and</strong> r<strong>and</strong>omized plot based sampling<br />

Rakoto Ratsimba et al. (2010) in their study on deforestation <strong>and</strong> degradation show that the<br />

differentiation between “low degraded forest” <strong>and</strong> “degraded forest” is possible in tropical<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H.R. Ratsimba et al. 2010. Multi-scale analysis of carbon stocks <strong>and</strong> deforestation monitoring<br />

400<br />

humid forest with SPOT 5 image. Their remote sensing analysis was combined with biomass<br />

inventory relating the difference between the carbon stocks between the two classes. It also<br />

demonstrates that this situation was possible at small scale analysis (SPOT images cover about<br />

60 km * 60 km area). However, it is not always possible to cover the whole country with SPOT<br />

image (problem of cost) for the change monitoring. In this way, this research is dealing with<br />

lower resolution image in the South East part of Madagascar as test region (see Figure 1):<br />

- a preclassification is realized on a L<strong>and</strong>ast-7 ETM+ 2005 image to have the different forest<br />

type according an a priori classification : “low degraded forest” <strong>and</strong> “degraded forest”. The<br />

other l<strong>and</strong> use is only put in a “non forest” class including the different secondary formations,<br />

- a biomass inventory through Rakoto Ratsimba et al. (2010) methods is realized with a local<br />

allometric equation establishment (through tree based sampling <strong>and</strong> wood density analysis) <strong>and</strong><br />

a r<strong>and</strong>omized plot sampling (see Figure 1),<br />

Figure 1: Test site (left) <strong>and</strong> distribution of the biomass inventory plots (right)<br />

L<strong>and</strong>sat 7 ETM+ 2005, composite image RGB: 4 – 2 – 1<br />

Projection: Laborde (Hotine Mercator Oblique) – Ellipsoïde International 1924<br />

- a correlation between the biomass plot stock <strong>and</strong> value of vegetation indices is set up. In fact,<br />

this correlation is not possible with L<strong>and</strong>sat image because of sensor problems (black pixels<br />

throughout the image). Thus, Normalized Difference Vegetation Index (NDVI) combined with<br />

the Enhanced Vegetation Index (EVI) of Moderate Resolution Imaging Spectroradiometer<br />

(MODIS) Image (see Figure 2) are derived from the same period of the biomass inventory<br />

(2009). The Figure 2 summarizes the research approach.<br />

3. Result<br />

3.1. <strong>Forest</strong> biomass stock variation<br />

The assessment of carbon stocks in the field shows that due to the accuracy of the stratification,<br />

the biomass stock is very low (4 t/ha) but can reach over 500 t/ha (at plot scale). It shows that<br />

the above-ground biomass estimate uncertainty is attributed to stratification error in this phase<br />

of the methodology. It is related to the spatial resolution of L<strong>and</strong>sat <strong>and</strong> the time period<br />

difference between the image (2005) <strong>and</strong> the biomass inventory (2009). In fact, the distribution<br />

of the points in the Figure 4 illustrates that the vegetation indices have not significant variations<br />

<strong>and</strong> gives the default value for the classification of NDVI <strong>and</strong> or EVI image.<br />

Thus, assessing above-ground biomass using vegetation indices derived from MODIS image<br />

proved that monitoring carbon at large scale is possible. The Figure 5 shows the map derived<br />

from the combined classification from EVI <strong>and</strong> NDVI. Data from the fields indicate that this<br />

map is corresponding to forest that has 126 ± 12 t/ha stocks of carbon.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H.R. Ratsimba et al. 2010. Multi-scale analysis of carbon stocks <strong>and</strong> deforestation monitoring<br />

401<br />

Figure 2: (1) MODIS (left) mosaic composite image RGB 2009: 3 – 2 – 1 (2)<br />

derived EVI image (up right) <strong>and</strong> (3) derived NDVI image (down right)<br />

Projection: Laborde (Hotine Mercator Oblique) – Ellipsoïde International 1924<br />

Spatial analysis<br />

Biomass inventory<br />

Allometric equation<br />

L<strong>and</strong>sat-7 ETM+ 2005<br />

Preclassification<br />

Modis EVI / NDVI<br />

Stratification<br />

Plot based sampling<br />

Tree based sampling<br />

EVI/NDVI values for classification<br />

Post classification: Biomass map<br />

Plot based biomass<br />

stock<br />

Wood density / weight<br />

Allometric equation<br />

Accuracy assessment<br />

Figure 3: Research Approach<br />

Vegetation indices (*1000)<br />

1000<br />

Biomass (t/ha)<br />

600<br />

750<br />

400<br />

500<br />

250<br />

200<br />

0<br />

0<br />

1 51<br />

Biomass (t/ha)<br />

101 151<br />

NDVI*1000<br />

201 251<br />

EVI*1000<br />

301 351<br />

Observations<br />

Figure 4: Repartition of NDVI <strong>and</strong> EVI with plot biomass stock<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H.R. Ratsimba et al. 2010. Multi-scale analysis of carbon stocks <strong>and</strong> deforestation monitoring<br />

402<br />

4. Discussion<br />

Although deforestation occurs in practically all developing countries (FAO, 2006), the actual<br />

rate of deforestation makes REDD initiatives more <strong>and</strong> more important to maximize reductions<br />

of greenhouse gases. In this way, it is important to create methodologies to assess <strong>and</strong> to<br />

monitor the deforestation <strong>and</strong> the carbon stock together. The REDD initiatives suppose that<br />

countries has to be able to assess both degradation <strong>and</strong> deforestation processes. However, the<br />

degradation analysis seems to have some limitations regarding the cost effectiveness of the<br />

methodology. Rakoto Ratsimba et al. (2010) have shown that this kind of assessment is possible<br />

at a small scale (local level) while the upscaling process demonstrates that only deforestation<br />

monitoring is possible over large areas. It is due to loss of spatial accuracy (the diversity of<br />

value of 23 * 23 pixels of SPOT image are included in a single pixel of MODIS image).<br />

This study reports that it is possible to report a forest map through biomass stock assessment. It<br />

gives the opportunity for policy makers to have not only the total area of the forest (2 754 905<br />

ha for the humid forest of the Eastern of Madagascar), but also the mean carbon stock per area<br />

(126 ± 12 t/ha) (see Figure 5). If we refer to other studies done in the region, Sussman et al.<br />

(1994) estimated the forest area of the region at 3 800 000 ha in 1985. Moreover, the same<br />

assessment can be done in other images (2000 for example) in order to model the deforestation<br />

process. The accuracy assessment gives a total value of 28,85 % which is an acceptable value<br />

regarding other carbon assessment (Samalca et al., 2007 – 41,27 % ; Rakoto Ratsimba et al.,<br />

2010 – 31,35 %). Thus, depending on REDD objective in Madagascar <strong>and</strong> the country priority, a<br />

small scale or larger scale initiative could be implemented.<br />

Figure 5: <strong>Forest</strong> map of the Eastern Humid <strong>Forest</strong> of Madagascar (biomass stock 126 ± 12 t/ha)<br />

Projection: Laborde (Hotine Mercator Oblique) – Ellipsoïde International 1924<br />

References<br />

Aboal J.R., Arevalo, J.R. <strong>and</strong> Fern<strong>and</strong>ez, A., 2005. Allometric relationships of different tree<br />

species <strong>and</strong> st<strong>and</strong> above ground biomass in the Gomera laurel forest (Canary Isl<strong>and</strong>s).<br />

Flora - Morphology, Distribution, Functional Ecology of Plants, 200(3): 264-274.<br />

Food <strong>and</strong> Agriculture Organization of the United Nations (FAO), 2006. <strong>Forest</strong> Resources<br />

Assessment. Rome.<br />

Bonan G. B., 2008. <strong>Forest</strong>s <strong>and</strong> Climate <strong>Change</strong>: Forcings, Feedbacks, <strong>and</strong> the Climate<br />

Benefits of <strong>Forest</strong>s. Science, 2008, vol. 320, no. 5882. pp. 1444.<br />

Central Intelligence Agency (CIA), 2009. The World Factbook 2009. Washington, DC.<br />

Holmgren P., Clairs T. <strong>and</strong> Kasten T., 2008. Role of satellite remore sensing in REDD. Issue<br />

Paper, UN REDD (Food <strong>and</strong> Agriculture Organization of the United Nations, United<br />

Nations Development Programme, United Nations Environment Programme), 9p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H.R. Ratsimba et al. 2010. Multi-scale analysis of carbon stocks <strong>and</strong> deforestation monitoring<br />

403<br />

Humbert H., 1927. La Destruction d'une Flore Insulaire par le Feu. Mémoires de L'Académie<br />

Malgache, Fascicule V, Tananarive.<br />

Intergovernmental Panel on Climate <strong>Change</strong> (IPCC), 2001. Climate <strong>Change</strong> 2001: Working<br />

Group I: The Scientific Basis. Cambridge University Press, New York.<br />

Intergovernmental Panel on Climate <strong>Change</strong> (IPCC), 2003. Good practice guidance for L<strong>and</strong><br />

Use, L<strong>and</strong> Use change <strong>and</strong> <strong>Forest</strong>ry (LULUCF). Institute for <strong>Global</strong> Environmental<br />

Strategies, Hayama, Japan.<br />

Intergovernmental Panel on Climate <strong>Change</strong> (IPCC), 2006. Guidelines for national greenhouse<br />

gas inventories. Vol. 4, Agriculture, <strong>Forest</strong>ry <strong>and</strong> other L<strong>and</strong> Use (AFOLU). Institute for<br />

<strong>Global</strong> Environmental Strategies, Hayama, Japan.<br />

Jenkins M. D., 1987. Madagascar: An Environmental Profile. IUCN, Gl<strong>and</strong>.<br />

Jolly A. <strong>and</strong> Jolly R., 1984. Malagasy economics <strong>and</strong> conservation: A tragedy without villains.<br />

In Jolly A., Oberlé P., <strong>and</strong> Rol<strong>and</strong> R. (eds.) Key Environments: Madagascar. Pergamon<br />

Press, Oxford, p. 211-217.<br />

Kraenzel M., Castillo A., Moore T. <strong>and</strong> Potvin C., 2003. Carbon storage of harvest-age teak<br />

(Tectona gr<strong>and</strong>is) plantations, Panama. <strong>Forest</strong> Ecology <strong>and</strong> Management, 173(1-3): 213-<br />

225.<br />

Laclau, P., 2003. Biomass <strong>and</strong> carbon sequestration of ponderosa pine plantations <strong>and</strong> native<br />

cypress forests in northwest Patagonia. <strong>Forest</strong> Ecology <strong>and</strong> Management, 180(1-3): 317-333.<br />

Lewis B.A., Phillipson P.B., Andrianarisata M., Rahajasoa G., Rakotomalaza P.J.,<br />

R<strong>and</strong>riambolona M. <strong>and</strong> McDonagh J.F., 1996. A study of the Botanical Structure,<br />

Composition, <strong>and</strong> Diversity of the Eastern Slopes of the Réserve Naturelle Intégrale<br />

d’Andringitra, Madagascar. In S. M. Goodman (ed). A floral <strong>and</strong> Faunal Inventory of the<br />

Eastern Slopes of the Réserve Naturelle Intégrale d’Andringitra, Madagascar: with<br />

reference to elevational variation. Fieldiana: Zoology, new series, 85: 24-75.<br />

Losi C.J., Siccama T.G., Condit R. <strong>and</strong> Morales J.E., 2003. Analysis of alternative methods for<br />

estimating carbon stock in young tropical plantations. <strong>Forest</strong> Ecology <strong>and</strong> Management,<br />

184(1-3): 355-368.<br />

McNeely J. A., Miller K. R., Reid W. V., Mittermeier R. A. <strong>and</strong> Werner T. B., 1990.<br />

Conserving the World's Biological Diversity. IUCN, WRI, CI, WWF-US, <strong>and</strong> World<br />

Bank, Gl<strong>and</strong> <strong>and</strong> Washington, D.C.<br />

Mittermeier R. A., Konstant W. R., Nicoll M. E., <strong>and</strong> Langar<strong>and</strong> O., 1992. Lemurs of<br />

Madagascar: An Action Plan for Their Conservation, 1993-1999. IUCN, Gl<strong>and</strong>.<br />

Myers, N., 1988. Threatened biotas: "Hotspots" in tropical forests. Environmentalist 8: 1-20.<br />

Population Reference Bureau (1992). World Populations Data Sheet. Washington, D. C.<br />

Rakoto Ratsimba H., Rajoelison L. G., Rakotondrasoa L. O., Eckert S., Hergarten C. et<br />

Ehrensperger A., 2010. Dégradation des forêts et stock de carbone dans la biomasse<br />

épigée de la forêt dense humide de Manompana – Nord Est de Madagascar. Conférence<br />

scientifique internationale sur les technologies en faveur du développement EPFL –<br />

Chaire UNESCO.<br />

Rauh W., 1979. Problems of biological conservation in Madagascar. In Bramweil D. (ed.),<br />

Plants <strong>and</strong> Isl<strong>and</strong>s. Academic press, London, p. 405-421.<br />

Samalca I. K., De Gier A. et Li Hussin Y., 2007. Estimation of tropical forest biomass for<br />

assessment of carbon sequestration using regression models <strong>and</strong> remote sensing in Berau,<br />

East Kalimantan, Indonesia. Department of Natural Resources. The International Institute<br />

for Geoinformation Science <strong>and</strong> Earth Observation (ITC). Asian Conference on Remote<br />

Sensing (ACRS). 6p.<br />

Sussman R. W., Richard A. F., <strong>and</strong> Ravelojaona G. (1985). Madagascar: Current projects <strong>and</strong><br />

problems in conservation. Primate Conservation 5: 53-59.<br />

Sussman R. W., Green G. M. <strong>and</strong> Sussman L. K., 1994. Satellite Imagery, Human Ecology,<br />

Anthropology, <strong>and</strong> Deforestation in Madagascar.Human Ecology, vol 22, N°3.<br />

White F., 1983. The vegetation of Africa, a descriptive memoire to accompany<br />

UNESCO/AETFAT Vegetation map of Africa. UNESCO, Paris.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Rodrigues et al. 2010. Characterization of a Maculinea alcon population in the Alvão Natural Park (Portugal)<br />

404<br />

Characterization of a Maculinea alcon population in the Alvão Natural<br />

Park (Portugal) by a mark-recapture method<br />

Maria Conceição Rodrigues 1,2 , Patrícia Soares 3 , José Aranha 1,2 & Paula Seixas<br />

Arnaldo 1,2*<br />

1<br />

Departamento de Ciências Florestais e Arquitectura Paisagista, Universidade de Trásos-Montes<br />

e Alto Douro, 5001-801 Vila Real, Portugal<br />

2<br />

CITAB, Universidade de Trás-os-Montes e Alto Douro, 5001-801 Vila Real, Portugal<br />

3<br />

GETER, Grupo de Estudos Territoriais, Universidade de Trás-os-Montes e Alto Douro,<br />

5001-801 Vila Real, Portugal<br />

Abstract<br />

The blue alcon, Maculinea alcon located at the Alvão Natural Park was studied by markrecapture<br />

methods in order to estimate population size <strong>and</strong> flight range of the butterflies.<br />

Sampling was made between 2007 <strong>and</strong> 2009. The results showed that maculinea population size<br />

has increased along the studied period with an estimated density at the peak of the flight period<br />

of 190 butterflies in 2007, 392 in 2008 <strong>and</strong> 1653 in 2009. The number of captured males was<br />

higher at the beginning of the flight period while females increase gradually over the flight<br />

period. The average value of the sex ratio was 0.9 in 2007, 1.4 in 2008 <strong>and</strong> 1.2 in 2009. The<br />

flight range of the butterflies did not show significant differences between sexes <strong>and</strong> an average<br />

value of about 12 m were recorded. Thus the results indicate that this population is in expansion<br />

<strong>and</strong> not threatened by extinction.<br />

Keywords: Maculinea alcon L, population estimate, mark-recapture method<br />

1. Introduction<br />

As a result of their extraordinary life cycle, Maculinea butterflies are typical representatives of<br />

the most threatened species in Europe (Wynhoff 1998; Mungira <strong>and</strong> Martin 1999; Van Swaay<br />

<strong>and</strong> Warren 1999). Like most other Lepidoptera species, Maculinea alcon lays the eggs on a<br />

selected host plant, the Gentiana pneumonanthe. However, caterpillars do not complete<br />

development on the plant. Instead they are voluntarily carried by Myrmica ants to their nests.<br />

Once in the nest, caterpillars are fed by the ants until complete larval development (Thomas<br />

1995). Thus, protecting the butterfly is protecting the whole complex systems because, both the<br />

host plant <strong>and</strong> the ant nests are essential for the successful development of Maculinea alcon. In<br />

Portugal, the Alvão Natural Park is the only known place where is it possible to see Maculinea<br />

alcon flying. In the past, several populations of the blue alcon were reported by local people but<br />

with the habitat fragmentation <strong>and</strong> ab<strong>and</strong>onment of extensive management many of those<br />

populations disappeared. However, there still exist some small populations that need to be<br />

studied. Therefore, the population density, the probability of survival <strong>and</strong> the dispersal ability<br />

are important study population features that can lead to programmes of conservation. In this<br />

study we examined population size, sex ratio <strong>and</strong> movements that characterize a Maculinea<br />

alcon population located at the Alvão Natural Park, which is a part of a Site of Community<br />

Importance (SCI) for the Mediterranean biogeographical region, listed in 2006 by the European<br />

Commission.<br />

* Corresponding author. Tel.: + 00 351 259 350 555<br />

Email address: parnaldo@utad.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Rodrigues et al. 2010. Characterization of a Maculinea alcon population in the Alvão Natural Park (Portugal)<br />

405<br />

2. Methodology<br />

The method of repeated marking-release-recapture was used during June, July <strong>and</strong> August<br />

between 2007 <strong>and</strong> 2009 in order to estimate the parameters characterizing the maculinea<br />

population. This method consists in the capture of as many individuals as possible. Each<br />

captured adult was sexed <strong>and</strong> market with a number on the underside of the right hind wing,<br />

using a black permanent pen. Immediately after that, the butterfly was released at the same point<br />

of capture. After three days, sampling was repeated <strong>and</strong> the basic parameters of population<br />

dynamics were estimated on the basis of the rate marked vs unmarked M. alcon captured.<br />

Number of sampling days varied along the studied years. Jolly-Seber method was used to<br />

estimate population size <strong>and</strong> survival probability on those years. In 2009, at each point of<br />

capture <strong>and</strong> recapture, GPS coordinates were recorded <strong>and</strong> used to calculate the flight range of<br />

the butterflies.<br />

3. Results<br />

In 2007, the capture took place between 11 to 31 July <strong>and</strong> a total of 265 butterflies were marked,<br />

154 males <strong>and</strong> 111 females. Of the market butterflies, 70 (26.4%) were recaptured only once<br />

<strong>and</strong> 8 (3.0%) were recaptured twice during the sampling period. The recapture rate was 29.4%.<br />

In 2008, a total of 263 were marked between 14 July <strong>and</strong> 7 August, 141 males <strong>and</strong> 122 females.<br />

Only 17 of the total market butterflies were recaptured (6.5%). In 2009, 566 butterflies were<br />

marked between 13 July <strong>and</strong> 10 August with 318 males <strong>and</strong> 248 females. Of the market<br />

butterflies, 43 (7.6%) butterflies were recaptured only once <strong>and</strong> 2 butterflies were recaptured<br />

twice. Sex ratio was calculated for each sampling day (see Figure 1). Results showed<br />

differences in the way males <strong>and</strong> females behave during the study years. The total number of<br />

captured males was higher than females during the studied years. However, the sex ratio varied<br />

across the sampling days. Although in 2007 the sex ratio was always higher than 1, favourable<br />

to males, the rate of females increased gradually which could be registered on the last two years<br />

with an inversion of the sex ratio at the middle of the flight period with more females after 24<br />

July. The result of the population size indicates that Maculinea alcon population has gradually<br />

increased along the study period (see Figure 2). In 2009, at the peak of the flight period, the<br />

estimated number of the butterflies was four times higher than in 2008 <strong>and</strong> even higher than in<br />

2007. The flight range of the butterflies measured by using capture <strong>and</strong> recapture GPS<br />

coordinates (see Figure 3) showed that from the 43 recaptured butterflies in 2009 only 14<br />

moved more than 0,05m (5 males <strong>and</strong> 9 females). On average, distances of about 11.7±3.2 m<br />

were recorded with maximum values of about 80 m with no significant differences between<br />

males <strong>and</strong> females flight range (F=0.13; gl=43; Sig=0.911).<br />

4. Discussion<br />

Results showed differences in population characteristics across the studied years. Although the<br />

beginning of the flight period were almost the same (second week of July), the end differed<br />

between years with a shorter flight period in 2007. This could be due to weather conditions<br />

which was very rainy in the beginning of August 2007. Also, differences were obtained for sex<br />

ratio <strong>and</strong> behaviour characteristics of males <strong>and</strong> females. In the last two years, the rate of<br />

females increased gradually <strong>and</strong> the approximately 1:1 sex ratio was observed in the last third of<br />

the flight period. This is important in the balance of the population because at the beginning of<br />

the flight period the host plant used by females to lay the eggs is not flowering yet <strong>and</strong> that can<br />

compromise the success of larvae development. Identical results were obtained by Árnyas et al<br />

(2005) in a Maculinea rebeli population in Hungary. The mobility results of the butterflies<br />

indicate that their dispersal ability is very low. According to Thomas (1991) this could be a<br />

serious problem for their survival in the modern European l<strong>and</strong>scape. The ability of dispersion<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Rodrigues et al. 2010. Characterization of a Maculinea alcon population in the Alvão Natural Park (Portugal)<br />

406<br />

is very important in the successional habitats that this species inhabits (Johnson 2000). When<br />

the habitat becomes unsuitable it is necessary migration movements <strong>and</strong> available suitable sites<br />

within very short distances are desirable<br />

References<br />

Árnyas, E., Bereczki, J., Tóth, A., Varga, Z., 2005. Results of the mark-release-recapture<br />

studies of a Maculinea rebeli population in the Aggtelek karst (N Hungary) between 2002-<br />

2004. In: Studies on the Ecology <strong>and</strong> Conservation of Butterflies in Europe. Vol2: Species<br />

ecology along a European gradient: buterflies as a model. Edited by J. Settele, E. Kuhn<br />

ans J. Thomas: 111-114.<br />

Johnson, M.P., 2000. The influence of patch demographics on metapopulations, with particular<br />

reference to successional l<strong>and</strong>scape. Oikos 88:67-74.<br />

Munguira, M.L. <strong>and</strong> Martin, J., 1999. Action Plan for Maculinea Butterflies in Europe. In:<br />

Council of Europe Publishing, editor. Convention on the Conservation of European<br />

Wildlife <strong>and</strong> Natural Habitats (Bern Convention), Nature <strong>and</strong> Environment, 97, Strasburg.<br />

Thomas, J.A., 1991. Rare species conservation: Case studies of European butterflies. In:<br />

Spellerberg, I.f., Goldsmith, F. B., Morris, M. G. (eds), The Scientific Management of<br />

Temperate Communities for Conservation. Blakwell Scientific Publications, Oxford: 149-<br />

197.<br />

Thomas, J.A., 1995. The ecology <strong>and</strong> conservation of Maculinea arion <strong>and</strong> other European<br />

species of large blue butterfly. In: A.S. Pullin, editor. Ecology <strong>and</strong> Conservation of<br />

Butterflies. London: Chapman & Hall. p. 180-197.<br />

Van Swaay, C.A.M. <strong>and</strong> Warren, M.S., 1999. Red Data Book of European Butterflies<br />

(Rhophalocera). Nature <strong>and</strong> Environment 99. Council of Europe Publishing, Strasbourg<br />

Wynhoff, I. 1998. The recent distribution of the European Maculinea species. J. Insect<br />

Conservation, 2: 15-27.<br />

Acknowledgement<br />

Authors would like to express their thanks to CITAB - UTAD who supported this work <strong>and</strong> a<br />

special gratitude to the Natural Park of Alvão <strong>and</strong> habitants of Lamas de Olo village for their<br />

contributions in our investigations.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Rodrigues et al. 2010. Characterization of a Maculinea alcon population in the Alvão Natural Park (Portugal)<br />

407<br />

2007<br />

sex ratio<br />

4<br />

3<br />

2<br />

1<br />

0<br />

06-Jul 11-Jul 16-Jul 21-Jul 26-Jul 31-Jul 05-Ago<br />

day<br />

2008<br />

2009<br />

sex ratio<br />

4<br />

3<br />

2<br />

1<br />

0<br />

11-Jul 16-Jul 21-Jul 26-Jul 31-Jul 05-Ago 10-Ago<br />

day<br />

sex ratio<br />

8<br />

6<br />

4<br />

2<br />

0<br />

11-Jul 16-Jul 21-Jul 26-Jul 31-Jul 05-Ago 10-Ago 15-Ago<br />

day<br />

Figure 1: Sex ratio for each sampling day in 2007, 2008 <strong>and</strong> 2009<br />

2007<br />

Estimated population (N)<br />

600<br />

400<br />

200<br />

0<br />

-200<br />

-400<br />

12 1,0<br />

136,5 15 8 ,1 82,3<br />

170,3<br />

0,0 9,3<br />

22,5<br />

0,0<br />

1 2 3 4 5 6 7 8 9<br />

Days<br />

2008<br />

2009<br />

Estimated population (N)<br />

2000,0<br />

1500,0<br />

1000,0<br />

500,0<br />

0,0<br />

-500,0<br />

-1000,0<br />

-1500,0<br />

35,3 19 ,0<br />

392,0<br />

0,0 36,0 48,0 31,5<br />

0,0<br />

1 2 3 4 5 6 7 8<br />

Days<br />

Estimated population (N)<br />

4000<br />

3000<br />

2000<br />

1652,8<br />

1000<br />

0<br />

153,6 60,8<br />

115,5 17 4 ,2 303,0<br />

36,0 0,0<br />

-1000 1 2 3 4 5 6 7 8 9<br />

Days<br />

Figure 2: Estimates of the number of individuals <strong>and</strong> st<strong>and</strong>ard error of estimation in 2007, 2008 <strong>and</strong> 2009<br />

by the Jolly-Seber method.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.C. Rodrigues et al. 2010. Characterization of a Maculinea alcon population in the Alvão Natural Park (Portugal)<br />

408<br />

Figure 3: Capture <strong>and</strong> recapture coordinates used to calculate dispersal ability of Maculinea alcon<br />

butterflies.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.T.C. Rodrigues & V. Silva 2010. L<strong>and</strong>scape changes in a watershed in the southwest of Portugal<br />

409<br />

L<strong>and</strong>scape changes in a watershed in the southwest of Portugal<br />

Maria Teresa Calvão Rodrigues * & Vanessa Silva<br />

Faculdade de Ciências e Tecnologia/Universidade Nova de Lisboa, Portugal<br />

Abstract<br />

Over the centuries human intervention has altered the l<strong>and</strong>scape of the south of Portugal, the<br />

most dramatic changes having taken place during the last one hundred <strong>and</strong> fifty years: the<br />

dominant component of the l<strong>and</strong>scape has changed first from natural/semi-natural shrubby<br />

communities to cereal crops <strong>and</strong> later to oak tree montados through processes of farml<strong>and</strong><br />

ab<strong>and</strong>onment, reforestation, change in the type of crops <strong>and</strong> trees grown, cutting or burning. All<br />

of these changes affected l<strong>and</strong>scape composition <strong>and</strong> configuration <strong>and</strong> thus the functioning of<br />

natural systems.<br />

The objective of this paper is to study l<strong>and</strong>scape dynamics of a watershed in the southwest of<br />

Portugal over a century, using data from different time periods <strong>and</strong> the consequences of<br />

changes. Within the context of a GIS l<strong>and</strong>-use changes were quantified through the analysis of<br />

l<strong>and</strong> use maps <strong>and</strong> ortophotomaps.<br />

Keyword: L<strong>and</strong> cover change, l<strong>and</strong>scape<br />

1.Introduction<br />

Human intervention significantly altered the l<strong>and</strong>scape of southern Portugal, the primitive<br />

forests having disappeared long ago. The l<strong>and</strong>scape of the South of the country suffered major<br />

changes in its structure <strong>and</strong> composition over the past 150 years. In the late nineteenth century<br />

the l<strong>and</strong>scape matrix essentially consisted of shrubl<strong>and</strong>s (Feio 1998). From that date on<br />

significant modifications that contributed decisively to soil degradation occurred. Successive<br />

campaigns to promote the production of cereals in the first half of the twentieth century led to a<br />

sharp increase of the cultivated area at the expense of farming on unsuitable l<strong>and</strong>: steep slopes,<br />

rocky soils, shallow <strong>and</strong> of low infiltration capacity. As Feio (1998) points out, within a time<br />

period of about 50 years the shrubl<strong>and</strong>s almost disappeared from the l<strong>and</strong>scape of southern<br />

Portugal while at the same time the plowed area with grain duplicated.<br />

However, crop cultivation on unsuitable l<strong>and</strong> led to intense erosion which in short time<br />

conditioned new sowings <strong>and</strong> consequently led to the ab<strong>and</strong>onment of the sites less apt for<br />

agriculture or its replacement by other forms of soil use. Thus, “Carta Agrícola e Florestal” for<br />

the decade 50-60, shows that in areas of steeper slopes, agriculture had been replaced by forests<br />

(montados) or left ab<strong>and</strong>oned, leading to a progressive recovery of the natural vegetation. In the<br />

70s there was a strong expansion of the eucalyptus culture. In most cases, the deployment of<br />

plantations was not well conducted, which once again led to soil degradation. From the 90s on a<br />

tendency for the ab<strong>and</strong>onment of cultivation of eucalyptus began to occur, or at least a lack of<br />

significant new investments due to the low economic profitability.<br />

* Corresponding author. Teresa Calvão Tel.: 212948397 - Fax: 212948554<br />

Email adress: mtr@fct.unl.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.T.C. Rodrigues & V. Silva 2010. L<strong>and</strong>scape changes in a watershed in the southwest of Portugal<br />

410<br />

2. Methodology<br />

2.1 Study area<br />

The area selected for this study corresponds to the watershed of Ribeiro do Canas, a sub-basin<br />

of the Sado River Basin (Fig. 1). It is small a watershed (5335 ha) that corresponds to about<br />

0.8% of the Sado Basin. The climate is mediterranean with average annual temperature being<br />

approximately 16.3 °C <strong>and</strong> annual precipitation around 575 mm.<br />

According to the biogeographic map of Costa et al. (2002) the territory in analysis is included in<br />

the Mediterranean West Iberian Province (Lusitan-Extremadurean Subprovince, Marianic-<br />

Monchiquensean Sector <strong>and</strong> Baixo Alentejano-Monchiquense Subsector). The head of the<br />

natural potential climatophillous vegetation series consists of the association Pyro<br />

bourgaeanae-Quercetum rotundifoliae that is, sclerophyllous forests dominated by Holm Oak<br />

(Quercus rotundifolia). The regressive succession consists of three stages. The first substitution<br />

stage of the forests (Hyacinthoido hispanicae-Quercetum cocciferae) is dominated by Kermes<br />

Oak (Quercus coccifera); the next regression stage consists of shrubl<strong>and</strong>s Retameto<br />

sphaerocarpae-Cytisetum bourgaei; a scrub community dominated by Gum Rockrose (Cistus<br />

ladanifer), the association Genisto hirsutae-Cistetum ladaniferi, represents the maximum<br />

degradation stage (Costa et al. 1998).<br />

The changes that occurred in part of the study area, namely alongside Ribeiro do Canas, can be<br />

observed on aerial photographs from 1947 (Fig. 2A) <strong>and</strong> 2004/05 (Fig. 2B). In 1947 even the<br />

most steep terrain was then cultivated with cereals. Only some scattered ork oak treas were left.<br />

It can also be observed that at that time riparian vegetation had been largely destroyed. In<br />

2004/05 a remarkable recovery not only of riparian vegetation but also of vegetation occupying<br />

the slopes can be noticed. An increase in biomass <strong>and</strong> structural diversity of plant communities<br />

is clearly evident. Nowadays the progressive succession has led to the appearance of tall shrub<br />

communites dominated by Kermes Oak, that is to say, the vegetation has almost reached the<br />

climax stage.<br />

2.2 Methods<br />

L<strong>and</strong> cover of the study area was characterized for four time periods: 1895, 1963, 1990 <strong>and</strong><br />

2004/05. L<strong>and</strong> cover was obtained by digitizing existing maps in analogical format for 1895<br />

(Carta Agrícola) <strong>and</strong> 1963 (Carta Agrícola e Florestal) <strong>and</strong> color orthophotomaps for 2004/05.<br />

L<strong>and</strong> cover for 1990 already existed in the form of a polygon shapefile (Carta de Ocupação do<br />

Solo). Eight L<strong>and</strong> Cover classes were considered: Agriculture, <strong>Forest</strong>, Semi-natural vegetation,<br />

Bare soil, Water, Urban, Eucalypt plantations <strong>and</strong> Olive st<strong>and</strong>s <strong>and</strong> Orchards. However, the last<br />

five classes had areal percentage values always fewer than 4% during the time period of<br />

analysis <strong>and</strong> will not be presented in this paper.<br />

A Digital Elevation Model was produced which allowed the determination of slope, solar<br />

radiation <strong>and</strong> topographic wetness index (Beven <strong>and</strong> Kirby 1979). The information from these<br />

biophysical parameters was intersected with the information of l<strong>and</strong> cover. All the analysis were<br />

performed in ArcGis 9.2 (ESRI) except the computation of solar radiation which was performed<br />

in ArcView 3.2 (ESRI) using the extension Solar Analyst (Fu <strong>and</strong> Rich 2002).<br />

3. Results<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.T.C. Rodrigues & V. Silva 2010. L<strong>and</strong>scape changes in a watershed in the southwest of Portugal<br />

411<br />

The changes in l<strong>and</strong> cover of the study area, over time, referring to the three most representative<br />

classes are presented in Figure 3. It can be observed that the study area witnessed the most<br />

important alterations between 1895 <strong>and</strong> 1963 while it changed less markedly from 1963 to<br />

2004/05. It may also be noticed that the "matrix" of the l<strong>and</strong>scape, which consisted of the class<br />

Semi-natural vegetation in 1895, changed to <strong>Forest</strong> thereafter.<br />

The Semi-natural vegetation in 1895 occupied most of the watershed (65%), having suffered a<br />

dramatic decline between 1895 <strong>and</strong> 1963. Henceforth this l<strong>and</strong> cover class did not undergo<br />

considerable changes, constituting about 5% of the study area. As regards Agriculture, this class<br />

occupied then a very low percentage of the area (about 2%) <strong>and</strong>, within 68 years, it increased<br />

substantially, reaching 29%. From 1963 to 2004/05 this class knew a decrease in its<br />

representativeness, more significant between 1990 <strong>and</strong> 2004/05. In the latter date the<br />

agricultural areas made up about 17% of the total area. The class <strong>Forest</strong> had a marked increase<br />

between 1895 <strong>and</strong> 1963, an insignificant decrease from 1963 to 1990 <strong>and</strong> a considerable<br />

increase between 1990 <strong>and</strong> 2004/05.<br />

In Figures 4, 5 <strong>and</strong> 6 an analysis of l<strong>and</strong> cover by biophysical parameters is presented.<br />

Agriculture in general was practiced in less steep terrain. By contrast, semi-natural vegetation,<br />

except in 1895, occupied areas with more pronounced slope. The <strong>Forest</strong>s were always located in<br />

all slope classes. In 1990 <strong>and</strong> 2004/05 semi-natural vegetation occupies, to a large extent, areas<br />

with high soil moisture content. Overall, agriculture is not practiced in areas with high soil<br />

moisture. In general semi-natural vegetation occurs in areas receiving low radiation.<br />

4. Discussion<br />

The l<strong>and</strong>scape transformations described in the present study are consistent with those that<br />

occurred in the south of the country, portrayed in various studies by other authors. After<br />

agriculture ab<strong>and</strong>onment in the less suitable areas shrub cover increased. Semi-natural<br />

vegetation at present occupies areas that receive a smaller amount of solar radiation <strong>and</strong> to a<br />

considerable extent areas of high moisture availability. That is to say, vegetation has mainly<br />

recovered along the steep slopes of the rivers. These communities are structurally compex <strong>and</strong><br />

consist of tall shrubs typical of the potential vegetation. This aspect has important consequences<br />

at the l<strong>and</strong>scape level as these communities may constitute ecological corridors <strong>and</strong> thus allow<br />

the increase of diversity.<br />

References<br />

Beven, K. J. <strong>and</strong> Kirby, M. J., 1979. A physically based, variable contributing area model for<br />

basin hydrology. Hydrological. Sciences Bulletin, 24: 43-69.<br />

Costa, J. C., Aguiar, C., Capelo, J. H., Lousã, M. <strong>and</strong> Neto, C., 1998. Biogeografia de Portugal<br />

Continental. Quercetea, 0: 5-56.<br />

Costa, J. C., Capelo, J., Espírito Santo, M. D. <strong>and</strong> Lousã, M., 2002. Aditamentos à vegetação do<br />

Sector Divisório Português. Silva Lusitana, 10: 199-128.<br />

Feio, M., 1998. A Evolução da Agricultura no Alentejo Meridional. As Cartas Agrícolas de G.<br />

Pery. As Difíceis Perspectivas Actuais na Comunidade Europeia. Edições Colibri, Lisboa,<br />

112 p.<br />

Fu, P. <strong>and</strong> Rich, P.M., 2002. A geometric solar radiation model with applications in agriculture<br />

<strong>and</strong> forestry. Computers <strong>and</strong> Electronics in Agriculture, 37: 25-35.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.T.C. Rodrigues & V. Silva 2010. L<strong>and</strong>scape changes in a watershed in the southwest of Portugal<br />

412<br />

Study area<br />

Sado river basin<br />

Figure 1: Location of the study area<br />

B<br />

Figure 2: Part of the study area in 1947 (A) <strong>and</strong> in 2004/05 (B)<br />

80<br />

70<br />

Area (%)<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

<strong>Forest</strong><br />

Agriculture<br />

Semi‐natural<br />

vegetation<br />

0<br />

1895 1963 1990 2004/05<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.T.C. Rodrigues & V. Silva 2010. L<strong>and</strong>scape changes in a watershed in the southwest of Portugal<br />

413<br />

Figure 3: L<strong>and</strong> cover changes of the three most representative l<strong>and</strong> use cover classes.<br />

Area (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

1895<br />

0-5%<br />

5-8%<br />

8-12%<br />

12-15%<br />

15-25%<br />

>25%<br />

Area (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

1963<br />

0-5%<br />

5-8%<br />

8-12%<br />

12-15%<br />

15-25%<br />

>25%<br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

Area (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

1990<br />

0-5%<br />

5-8%<br />

8-12%<br />

12-15%<br />

15-25%<br />

>25%<br />

Area (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

2004/05<br />

0-5%<br />

5-8%<br />

8-12%<br />

12-15%<br />

15-25%<br />

>25%<br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

Figure 4: Slope values by l<strong>and</strong> cover class.<br />

100<br />

80<br />

Extremely dry<br />

Ve ry dry<br />

Dry<br />

1895<br />

100<br />

80<br />

Extremely dry<br />

Ve ry dry<br />

Dry<br />

1963<br />

Area (%)<br />

60<br />

40<br />

Moderate<br />

We t<br />

Ve ry we t<br />

Area (%)<br />

60<br />

40<br />

Moderate<br />

We t<br />

Ve ry we t<br />

20<br />

20<br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

Area (%)<br />

100<br />

80<br />

60<br />

40<br />

Extremely dry<br />

Ve ry dry<br />

Dry<br />

Moderate<br />

We t<br />

Ve ry we t<br />

1990<br />

Area (%)<br />

100<br />

80<br />

60<br />

40<br />

Extremely dry<br />

Ve ry dry<br />

Dry<br />

Moderate<br />

We t<br />

Ve ry we t<br />

2004/05<br />

20<br />

20<br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

Figure 5: Topographic wetness index by l<strong>and</strong> cover class.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.T.C. Rodrigues & V. Silva 2010. L<strong>and</strong>scape changes in a watershed in the southwest of Portugal<br />

414<br />

100<br />

80<br />

Ve ry lo w R a dia tio n<br />

Lo w R a dia tio n<br />

Medium Radiatio n<br />

1895<br />

100<br />

80<br />

Ve ry lo w R a dia tio n<br />

Lo w R a dia tio n<br />

Medium Radiation<br />

1963<br />

Area (%)<br />

60<br />

40<br />

High radiation<br />

Very high Radiation<br />

Area (%)<br />

60<br />

40<br />

High radiatio n<br />

Very high Radiation<br />

20<br />

20<br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

100<br />

80<br />

Ve ry lo w R a dia tio n<br />

Lo w R a dia tio n<br />

Medium Radiation<br />

1990<br />

100<br />

80<br />

Ve ry lo w R a dia tio n<br />

Lo w R a dia tio n<br />

Medium Radiation<br />

2004/05<br />

Area (%)<br />

60<br />

40<br />

High radiation<br />

Very high Radiation<br />

Area (%)<br />

60<br />

40<br />

High radiation<br />

Very high Radiation<br />

20<br />

20<br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

0<br />

Agriculture Semi-natural vegetation <strong>Forest</strong><br />

Figure 6: Solar radiation by l<strong>and</strong> cover class.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Schaefer-Santos & C. Lingnau 2010. <strong>L<strong>and</strong>scapes</strong> in Transition – Monitoring in Areas of L<strong>and</strong>slides<br />

415<br />

<strong>L<strong>and</strong>scapes</strong> in Transition – Monitoring in Areas of L<strong>and</strong>slides<br />

Jorgeane Schaefer-Santos * & Christel Lingnau<br />

Curitiba Federal University, Brazil<br />

Abstract<br />

The satellite images have been very effective for monitoring the l<strong>and</strong>scapes dynamics.<br />

<strong>L<strong>and</strong>scapes</strong> vulnerable to disasters can be monitored by change detection techniques. We<br />

applied these techniques in areas affected by l<strong>and</strong>slides in November 2008 in Morro do Bau,<br />

Santa Catarina State, which led to material <strong>and</strong> human losses. The study used four images from<br />

different dates between 1992 <strong>and</strong> 2009. Vegetation index were developed using b<strong>and</strong>s 7 <strong>and</strong> 4,<br />

minimizing the atmospheric <strong>and</strong> radiometric effects. The techniques used were effective to<br />

detect changes caused by the disaster, such as soil exposure <strong>and</strong> deposition; during the study<br />

period there was no deforestation in the area of l<strong>and</strong>slides; the event had magnitude greater than<br />

the forests protection could promote, mainly high rainfall, combined with the high potential<br />

erosion of these mountains sediments. These l<strong>and</strong>scapes are subject to natural relief<br />

accommodation events that will inevitably be disaster if associated with human occupation.<br />

Keywords: change detection techniques; l<strong>and</strong>slides; environmental disaster; Itajaí Valley;<br />

l<strong>and</strong>scape dynamic.<br />

1. Introduction<br />

The Itajaí Basin covers an area economically very important for the State of Santa Catarina,<br />

southern Brazil. The mountainous relief prints fragility of l<strong>and</strong>slides exacerbated by other<br />

regional environmental problems as water pollution, illegal occupations <strong>and</strong> sewer without<br />

treatment. Over time, with the rural exodus <strong>and</strong> urban growth, the region became even more<br />

subject to problems of burial <strong>and</strong> flooding produced by inadequate occupation of urban l<strong>and</strong> <strong>and</strong><br />

silting <strong>and</strong> contamination results in erosion <strong>and</strong> lack of rural waste treatment <strong>and</strong> urban<br />

(Schaefer-Santos, 2003).<br />

All these changes in l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> use can be detected through time series of satellite<br />

images with the application of techniques of change detection. The mapping can be<br />

accomplished from the comparison of classified images, subtraction of b<strong>and</strong>s of the same area at<br />

different times, vegetation indices, principal component analysis <strong>and</strong> neural networks (Carvalho<br />

Jr & Silva, 2007).<br />

Some changes, such as the increase or decrease in forest cover, can be evaluated based on<br />

vegetation index. Caring for the use of vegetation index should be taken, especially the false<br />

changes that may be related to phenological changes, which are those that occur with various<br />

species, such as leaf drop, flowering time, appearance of new leaves in spring (Matínez &<br />

Gilabert, 2009). In a long term, these changes found may be related to climate change <strong>and</strong> largescale<br />

disturbances antrogênicos (Bradley et al., 2007).<br />

To be chosen images from similar periods of the year, representing the same seasonal cycle of<br />

the forest, the effects of phenological changes are minimized as the effects generated by the<br />

shadow of a region with rugged. Janoth et al. (2007) say that images L<strong>and</strong>sat–TM <strong>and</strong> SPOT XI<br />

* Corresponding author. Tel.: 55 41 3360-4218 - Fax: 55 41 3360 4211<br />

Email address: eng_jorgeane@yahoo.com.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Schaefer-Santos & C. Lingnau 2010. <strong>L<strong>and</strong>scapes</strong> in Transition – Monitoring in Areas of L<strong>and</strong>slides<br />

416<br />

showed reliable results for detecting changes in forest areas damaged by wind <strong>and</strong> snow sliding<br />

snow in Sweden <strong>and</strong> Finl<strong>and</strong>. The procedures used by the authors based on the simple<br />

difference of multitemporal images of the spectral b<strong>and</strong>s of red <strong>and</strong> of mid infrared <strong>and</strong> change<br />

detection pixel to pixel <strong>and</strong> neighborly relations. The mid infrared, especially, pointed to change<br />

in areas of climax forest in which there was the removal of timber, <strong>and</strong> therefore this technique<br />

very effective for monitoring forest. In this study we sought, through the vegetation index found<br />

that deforestation could be a predominant factor in l<strong>and</strong>slides occurred in the study area in<br />

November 2008 which caused the death of 135 people for burial <strong>and</strong> for such were used images<br />

L<strong>and</strong>sat 5 TM.<br />

2. Methodology<br />

The study area is located in the Itajaí Basin, which lies between latitudes 26º20’ e 27º50’ e<br />

longitudes 48º40’ e 50º20’, extending for 150 km north to south <strong>and</strong> 155 km from east to west<br />

(Gaplan, 1988). The study area is located from the middle course of the river Itajaí, <strong>and</strong> it parts<br />

of the Middle subbacia Itajaí <strong>and</strong> part of Rio subbacia Luis Alves, encompassing the areas of<br />

disaster in 2008 (Figure 1). In the watershed area of the Rio Itajaí various formations are found<br />

among the Dense Atlantic Rain <strong>Forest</strong> <strong>and</strong> some nuclei of Araucaria <strong>Forest</strong> in the high Itajaí<br />

Valley. In the middle <strong>and</strong> lower regions of Itajai Valley, there is the Dense Atlantic Rain<br />

<strong>Forest</strong>, especially characterized by high density <strong>and</strong> high species richness of trees, saplings <strong>and</strong><br />

shrubs, <strong>and</strong> high density of epiphytes. (Klein, 1978).<br />

The study used 4 images as mentioned below:<br />

1. 10/6/1992;<br />

2. 30/6/1999;<br />

3. 4/6/2007;<br />

4. 1/2/2009.<br />

The images were so selected because they are of the same month (June) acquisition, except the<br />

last. The choice of images from the same month, minimized the differences in reflectance due to<br />

the effects of phenology of forest <strong>and</strong> shade relief. The relief shading in images acquired in<br />

similar periods of the year, for example, in the same month, the shaded areas are virtually<br />

coincident, due to the similarity of the angle of inclination <strong>and</strong> azimuth of the sun. The image<br />

had to be t4 February, for not having another cloudless after the environmental disaster.<br />

Figura 1: Study area localization.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Schaefer-Santos & C. Lingnau 2010. <strong>L<strong>and</strong>scapes</strong> in Transition – Monitoring in Areas of L<strong>and</strong>slides<br />

417<br />

2.1 <strong>Change</strong> Detection<br />

The purpose of applying the techniques of change detection is within a set of pixels, identifying<br />

those that are significantly different between the first <strong>and</strong> last image in sequence (Radke et al.<br />

2005). Select the image start (t1) <strong>and</strong> the final image (t2), where the procedure is repeated<br />

between pairs “t 2 – t 3 ” e “t 4 – t 3 ” , generating three classified images (<strong>Change</strong> Detection 1 – CD 1 ,<br />

<strong>Change</strong> Detection 2 – CD 2 e <strong>Change</strong> Detection 2 – CD 3 ). The classification of this study followed<br />

values change between -5 <strong>and</strong> +5. <strong>Change</strong>s close to zero (from -2 to +2) were ignored,<br />

representing "no change", <strong>and</strong> "abrupt change" were identified as increased <strong>and</strong> reduced<br />

biomass. We used the simple difference method <strong>and</strong> the percentage difference between<br />

Vegetation Index on four occasions (t1, t2, t3 <strong>and</strong> t4) as follows (1):<br />

D( x)<br />

= t2 ( x)<br />

− t1(<br />

x)<br />

(1)<br />

Onde:<br />

D(x) = classified image by simple difference;<br />

t 2 (x) = vegetation index in the final time of the analysis;<br />

t 1 (x) = vegetation index in the inicital time of analysis.<br />

3. Results<br />

In the study area can easily locate the areas of l<strong>and</strong>slides <strong>and</strong> areas with sediment deposition by<br />

staining differentiated image 4 (2009), with values close to zero, which represents a lack of<br />

vegetation. In this same place, it was observed that the vegetation index have no color<br />

differentiation mean that deforestation in the years before the l<strong>and</strong>slide (Figure 2). Figure 3<br />

shows the image obtained of the changes between 2007 <strong>and</strong> 2009 (CD 3) which can be observed<br />

in the l<strong>and</strong>slide area <strong>and</strong> increasing the forest cover.<br />

Figura 2: <strong>Change</strong> detection occurred in the area of l<strong>and</strong>slides since 1992 to 2009.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Schaefer-Santos & C. Lingnau 2010. <strong>L<strong>and</strong>scapes</strong> in Transition – Monitoring in Areas of L<strong>and</strong>slides<br />

418<br />

Figura 3: <strong>Change</strong> detection between years 2007 <strong>and</strong> 2009.<br />

4. Discussion<br />

The l<strong>and</strong>scape had some changes during the study period. Over time, the study area showed an<br />

increase in forest cover (Schaefer-Santos, 2003). In this study we observed that between 1992<br />

<strong>and</strong> 1999 (CD 1), there were small increases in forest cover <strong>and</strong> in some places, small losses<br />

(Figure 2). Moreover, it is observed between 1999 <strong>and</strong> 2007 (CD 2) that there is greater<br />

movement in the l<strong>and</strong>scape, leading to an increase forest cover in general, since the<br />

displacement curve to the area gain of vegetation. Already between 2007 <strong>and</strong> 2009 (CD 3), as<br />

shown in Figure 2 <strong>and</strong> detailed in Figure 3 is possible to identify the great movement that<br />

occurred in the l<strong>and</strong>scape, as areas of l<strong>and</strong>slides are very clearly identified, ie, loss the area<br />

forests caused by l<strong>and</strong>slides. These l<strong>and</strong>slides clearly seen on L<strong>and</strong>sat TM 5 may be considered<br />

high intensity occurred in as many areas of surfaces greater than 800m in length. Among the<br />

largest, most convergence was associated with water, especially ending in rivers. In this study<br />

area, population density is low, however, the weathering mantles are very deep <strong>and</strong> very fine<br />

sediments associated with the relief <strong>and</strong> deep valleys (Gaplan, 1988), which receive large<br />

amounts of marine moisture leading to a natural process of denudation. These circumstances<br />

lead to the natural conditions for l<strong>and</strong>slides. Small human changes that have occurred were not<br />

detected in this analysis. These possible anthropogenic changes certainly added to the natural<br />

conditions may have led to an increased risk factor for disasters. However, the masses of soil<br />

were soaked for 04 months due to heavy rains, which were associated with a weather system<br />

that meant that only between 21st to 22nd November 2008, occurred rainfall of 300 mm of rain<br />

in the Valley Itajaí, with accumulated of 1000 mm in November, the normal precipitation is<br />

1.800 mm in a year (Dias, 2009). This precipitation was two times greater than the rainfall<br />

occurred in April 2010 in Rio de Janeiro, however, due to population density, in Rio de Janeiro<br />

there were about 250 deaths, as in Santa Catarina, the number was only 135 people killed due to<br />

l<strong>and</strong>slides. Natural factors associated with high precipitation were prevalent for the occurrence<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Schaefer-Santos & C. Lingnau 2010. <strong>L<strong>and</strong>scapes</strong> in Transition – Monitoring in Areas of L<strong>and</strong>slides<br />

419<br />

of l<strong>and</strong>slides in Santa Catarina in November 2008, leading one to believe that these events are<br />

natural <strong>and</strong> are part of the natural transformation of the l<strong>and</strong>scape.<br />

References<br />

Bradley, B.A; JACOB, R. W.; Hermance, J. F. & Mustard, J. F. A curve fitting procedure to<br />

derive inter-annual phonologies from time series of noisy satellite NDVI data. In: Remote<br />

Sensing of Environment (2007). doi:10.1016/j.rse.2006.08.002.<br />

Carvalho JR, O. A. & Silva, N. C. da. Detecção de Mudança Espectral uma nova metodologia<br />

para análise de séries temporais. In: Anais XIII Simpósio Brasileiro de Sensoriamento<br />

Remoto, Florianópolis, Brasil, 21-26 abril 2007, INPE, p. 5635-5641.<br />

Dias, A. F. S. As chuvas de novembro de 2008 em Santa Catarina: um estudo de caso vis<strong>and</strong>o<br />

a melhoria do monitoramento e da previsão de eventos extremos. Dias, A. F. S (editor<br />

responsável). Nota Técnica do IPNE, INMET, EPAGRI. Disponível em meio digital:<br />

www.ciram.com.br/ciram _arquivos/arquivos/gtc/downloads/NotaTecnica_SC.pdf<br />

GAPLAN – Gabinete de Planejamento do Estado de Santa Catarina. Atlas de Santa Catarina.<br />

Rio de Janeiro: Aerofoto Cruzeiro. 1986. 173p.<br />

Janoth, J.; Eisl, M.; Klaushofer, F. & Luckel, W. Procedimentos baseados em segmentação para<br />

análise de mudanças e classificação florestais com dados de alta resolução. In:<br />

BLASCHKE, T& KUX, H. (Org.) Sensoriamento Remoto e SIG Avançados: novos<br />

sistemas sensores e métodos inovadores / versão brasileira; tradução de Hermann Kux. –<br />

2ª ed. – São Paulo: Oficina de Textos, 2007. p. 99 a 107.<br />

Klein, R.M. Mapa Fitogeográfico de Estado de Santa Catarina. Itajaí: Herbário Barbosa<br />

Rodrigues, 1978. Escala 1:1.000.000. 24p.<br />

Martinez, B. & Gilabert, M. A. Vegetation dynamics from NDVI times series analysis using the<br />

wavelet transform. In: Remote Sensing of Environment (2009). doi:<br />

10.1016/j.rse.2009.04016.<br />

Radke, R. J., Al-Kofahi, O. & Roysam, B. Image <strong>Change</strong> Detection Algorithms: A Systematic<br />

Survey. In: IEEE Transaction on Geoscience <strong>and</strong> Remote Sensing, vol. 14, n° 03. March<br />

2005. doi: 10.1109/TIP.2004.838698.<br />

SCHAEFER-SANTOS, J. Ocupação do solo e comportamento hidrológico da sub-bacia do Rio<br />

Luis Alves, bacia do Rio Itajaí, Santa Catarina. Dissertação de Mestrado, Curso de Pós-<br />

Graduação em Engenharia Florestal, UFPR: Curitiba, PR. 199p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Scheffer & E. Schimanski 2010. The process of urbanization <strong>and</strong> the environment<br />

420<br />

The process of urbanization <strong>and</strong> the environment - the case of<br />

irregular occupations in the city of Ponta Grossa, PR, Brazil<br />

S<strong>and</strong>ra Maria Scheffer & Édina Schimanski<br />

State University of Ponta Grossa, Brazil<br />

Abstract<br />

This article analyzes the process of urbanization as a social <strong>and</strong> historical construction,<br />

<strong>and</strong> housing problems arising from this process, through the installation of homes in<br />

illegal occupations in the city of Ponta Grossa - Paraná. Trying to underst<strong>and</strong> the<br />

dynamics of this phenomenon it explores the socio spatial location as the process of<br />

degradation of the environment, researching the specifics that arise for public policy.<br />

The methods used were a literature search <strong>and</strong> document research-based Plan for Social<br />

Housing. Thus, it identifies the planning of urban space by a team of professionals from<br />

diverse fields as fundamental to promote alternative solutions with adequate<br />

infrastructure <strong>and</strong> legality for the families as well as environmental improvement<br />

projects that provide improvements to support the inclusion social <strong>and</strong> sustainability.<br />

Keywords: urbanization, illegal occupations, environment<br />

1 - The process of urbanization<br />

The urbanization process is a dynamic process, arising from social relations in each<br />

historical context related changes in processes of social production <strong>and</strong> the manner of<br />

insertion of each society in general dynamics of capitalism. CASTELLS in his<br />

conception over the urban phenomenon means that the analysis of urbanization should<br />

follow the knowledge of its social process. It states that: "Explaining the social process<br />

that underlies the organization of space is not simply to place the urban phenomenon in<br />

context. A sociological problems of urbanization should consider it as a process of<br />

organization <strong>and</strong> development, <strong>and</strong> therefore, based on relationship between the<br />

productive forces, social classes <strong>and</strong> cultural forms (among which the space). " (1983,<br />

p.36)<br />

The human world changes <strong>and</strong> transformations are becoming more intense <strong>and</strong> fast.<br />

This is also reflected in the organization of space, which is being organized as diverse<br />

<strong>and</strong> complex. Many of their reorganizations made themselves <strong>and</strong> continue to happen,<br />

given the claims of production. Thus, the area is also historic, transforming itself<br />

through the social changes occurring over time.<br />

Urbanization has become a reality in Brazil, especially since the second half of the<br />

twentieth century, it was from the 1970s that the rural-urban ratio began to reverse the<br />

concentration of population to urban. The phenomenon of population concentration in<br />

urban areas has changed the dynamics of Brazilian society, resulting in a new lifestyle,<br />

urban style.<br />

With differences in each region <strong>and</strong> each municipality, the phenomenon of<br />

urbanization has spread throughout the national territory. Ponta Grossa, inl<strong>and</strong> city of<br />

Paraná State, was not immune to the phenomenon of urbanization. Although the Paraná<br />

be a state with a strong presence in agricultural production <strong>and</strong> Ponta Grossa have this<br />

characteristic was from the 1960s it became clear the population concentration in urban<br />

space.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Scheffer & E. Schimanski 2010. The process of urbanization <strong>and</strong> the environment<br />

421<br />

In the history of Ponta Grossa, you can verify outst<strong>and</strong>ing phases of development<br />

that contributed to the emergence <strong>and</strong> expansion of population <strong>and</strong> its urbanization<br />

process, as the occupation by the troopers, the construction of railroads accelerating<br />

process of urbanization, migration movements, especially foreigners, broadening the<br />

demographic structure, increasing industrialization through industrial development<br />

plans, among other factors.<br />

Among the most significant characteristics of the municipality, is its privileged<br />

geographical position, since its emergence as a way for troopers to the installation of<br />

major highways <strong>and</strong> railroads. It is considered a major road <strong>and</strong> rail junctions in the<br />

south of the country, due to its physical location <strong>and</strong> its access roads that allow the<br />

various regions of the state <strong>and</strong> interstate.<br />

The fact that they constitute a road-rail junction favored high density because it<br />

facilitated the transit of persons between the cities, especially those in nearby regions.<br />

In 2009, according to the IBGE, the municipality of Ponta Grossa had a population of<br />

314,681 inhabitants. Of the total population of the municipality, the rural end-grossense<br />

corresponds to 2.54% of total population <strong>and</strong> 97.46% of the population resides in urban<br />

area.<br />

These indices are above the national average, where 81.23% of the population is<br />

urban, which reflects on the one h<strong>and</strong> an accelerated process of urbanization <strong>and</strong> the<br />

other shows, that the city in recent decades had characterized the growth <strong>and</strong> / or<br />

maintenance a significant rural population.<br />

Thus, the urban space becomes the venue of expression of social struggles <strong>and</strong> conflicts<br />

of the capitalist order which generate dem<strong>and</strong> for the state, among them the dem<strong>and</strong> for<br />

housing.<br />

2 - The irregular occupations in Ponta Grossa <strong>and</strong> its socio-spatial<br />

The city resulting in the spatial form of the urbanization process has as one of its<br />

most striking characteristics, diversity. The city has its own dynamics, <strong>and</strong> in this<br />

context that the social needs arise more acutely among those to housing.<br />

The dynamics of urban social expresses itself in different forms of sociospatial<br />

structuration. Analyze how people try to solve their housing problem allows us to<br />

reflect on the factors that are engendered <strong>and</strong> which the contradictions that run through<br />

urban space.<br />

Urbanization as a social <strong>and</strong> historical process, it expresses the logic of the use<br />

<strong>and</strong> occupation of urban l<strong>and</strong>. The population dem<strong>and</strong>s the fulfillment of their needs to<br />

live in the urban environment. Not being met, suffer the consequences, among other<br />

factors, lack of infrastructure, competition for urban space being marked mainly by<br />

occupations public areas, which in the case of Ponta Grossa is facilitated by the relief<br />

which is very bumpy <strong>and</strong> full of streams.<br />

Lack of access to housing has led the impoverished segments of the urban<br />

population living in subhabitações so disorderly <strong>and</strong> without infrastructure, on l<strong>and</strong><br />

belonging to the government or unoccupied areas owned by individuals. This expresses<br />

the precariousness of insertion in the labor market, not from the incorporation of the<br />

entire workforce in the formal sector of the market, favoring the inclusion of a<br />

significant portion of the population in the informal sector <strong>and</strong> underemployment,<br />

accentuating social inequalities.<br />

In Ponta Grossa, the location of irregular occupation has a tendency to settle on the<br />

banks of streams in areas for permanent preservation, causing the destruction of riparian<br />

vegetation, thus altering the original l<strong>and</strong>scape. This tendency causes many problems<br />

for environmental preservation areas <strong>and</strong> for families living in these places because<br />

often suffer life-threatening situations <strong>and</strong> unsanitary.<br />

According to data from the Municipal Plan for Social Housing - 2010, Ponta<br />

Grossa has 8.778 housing units in the situation of illegal occupation in 162 points<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Scheffer & E. Schimanski 2010. The process of urbanization <strong>and</strong> the environment<br />

422<br />

distributed within the city limits, with a large percentage in the valleys, areas with no<br />

conditions for building <strong>and</strong> sanitation. On the other side also contribute to loss of<br />

biodiversity <strong>and</strong> / or change with the removal of riparian vegetation which results in<br />

pollution of local water contamination of streams, erosion of hillsides, l<strong>and</strong>slides, <strong>and</strong><br />

other factors that affect social development <strong>and</strong> conservation of natural habitat <strong>and</strong><br />

environment.<br />

To illustrate the situation, shown in Figure 1 to identify the illegal occupations in<br />

the urban area of Ponta Grossa.<br />

Figure 1 - Irregular occupations in the urban area of Ponta Grossa - 2010<br />

Source: Plan for Social Housing Ponta Grossa – 2010<br />

Some occupations are more recent, but others have already been established for<br />

decades in places. As Lowen SAHR (2001), the former slum in Ponta Grossa emerged<br />

in the 1950s, intensifying in the following decades, representing 0.8% of urban<br />

population in 1960 to 1.9% in 1970, 6.3% in 1980 14% in 1991<strong>and</strong> according to data<br />

from the Municipal Housing Plan, 11,03% in 2010 .<br />

Important to note that not all families living in irregular occupations live in<br />

conditions that can be considered slums because not every area of occupation reveals<br />

the precariousness of living conditions. Some occupations, even though in irregular<br />

areas, belonging to the municipality have a better quality of habitability <strong>and</strong><br />

environmental suitability.<br />

According to the Municipal Plan for Social Housing in the city of Ponta Grossa,<br />

we evaluated the inadequacy of environmental variable <strong>and</strong> for each illegal occupation<br />

of the urban area of Ponta Grossa was awarded points if you are on APP-Area<br />

Preservation <strong>and</strong> case is about risk area. The sum of two sub-components reflects the<br />

degree of inadequacy with regard to environmental conditions.<br />

Of the 162 points of irregular occupation, 80 points are in conditions considered<br />

high or very high environmental inadequacy, which corresponds to 49,4% of all<br />

occupations, alarming to propose coping strategies.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Scheffer & E. Schimanski 2010. The process of urbanization <strong>and</strong> the environment<br />

423<br />

3 - Environment <strong>and</strong> irregular occupations<br />

Analyze the relationship between urbanization <strong>and</strong> the environment is fundamental<br />

to underst<strong>and</strong> the mediations that arise between the need of social housing <strong>and</strong><br />

environmental respect.<br />

In the case of Ponta Grossa, its urban area has about only 150 km of streams in<br />

urban <strong>and</strong> largely due to its topography in the valleys, the irregular occupations that are<br />

located in these locations.<br />

The preservation of the environment is an essential theme <strong>and</strong> special Ponta<br />

Grossa, this ecological consciousness brings with it a need for a process of<br />

environmental education.<br />

Thus, planning the organization of urban space must be examined by a team of<br />

professionals in various areas in partnership with the communities concerned to<br />

promote alternative solutions with adequate infrastructure <strong>and</strong> legally for the families as<br />

well as environmental remediation projects that provide improvements to support social<br />

inclusion <strong>and</strong> sustainability.<br />

Thus, including the population living in irregular occupations, located in the<br />

vicinity of streams, in all steps to propose <strong>and</strong> implement improvements would create<br />

an interaction between nature <strong>and</strong> man <strong>and</strong> not treating them as distinct, but as<br />

interacting components in environment.<br />

As the prerogative of the Brazilian Agenda 21, sustainable development seeks to<br />

improve the lives of all people without increasing the use of natural resources beyond<br />

the capacity of the earth.<br />

Thus, the urban planning is crucial to the process of urbanization can be worked in<br />

order to avoid further problems <strong>and</strong> deterioration in physical space. With the relocation<br />

of families that generate environmental degradation of streams <strong>and</strong> riparian areas <strong>and</strong><br />

consequently live in areas at risk <strong>and</strong> secondly the regularization of lawfully made<br />

possible thereby, whether it was complying with an integrated planning in the city <strong>and</strong><br />

thinking of all points that comprise the process for solving problems arising from the<br />

socio historical process of urbanization in Ponta Grossa <strong>and</strong> consequent environmental<br />

degradation.<br />

It is also essential to undertake actions consistent to prevent new jobs mainly in<br />

environmental areas <strong>and</strong> providing housing programs to meet the housing law. With<br />

these proposals would preserve the care of social rights as constitutionally protected<br />

environmental law <strong>and</strong> housing rights, which have the same conceptual core, which is<br />

the socio-environmental function of property.<br />

The great challenge for the actors involved in the urban setting is to reconcile these<br />

two rights <strong>and</strong> build a scenario. It is necessary to adopt an inclusive model of urban<br />

growth where the humans can live in nature in a sustainable way in terms of<br />

promotions.<br />

References<br />

Castells, Manuel. The urban question. Translation of Arlene Caetano. Rio de Janeiro:<br />

Paz e Terra, 1983. 590 p.<br />

Brazilian Instituteof Geography <strong>and</strong> Statistics - IBGE. 2007 municipal population data.<br />

Rio de Janeiro: IBGE, 2001.<br />

Lowen Sahr, Ciciliano Luiza. Internal structure <strong>and</strong> social dynamics in the city of Ponta<br />

Grossa. In: DITZEL, Carmencita of Holleben Mello; LOWEN SAHR, Ciciliano<br />

Luiza (eds) Space <strong>and</strong> culture. Ponta Grossa <strong>and</strong> the Campos Gerais. Ponta<br />

Grossa: Editora UEPG, 2001. p. 13-36<br />

Oliveira, Isabel Eiras de. Status of the city, to underst<strong>and</strong> ... Rio de Janeiro: IBAM<br />

/DUMA, 2001. 64 p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.M. Scheffer & E. Schimanski 2010. The process of urbanization <strong>and</strong> the environment<br />

424<br />

Rolnik, Rachel; Nakano, Kazuo. Old questions, new challenges. In: Another urban<br />

world is possible: Notebooks Le Monde Diplomatique, Abaporu Institute, São<br />

Paulo, n. 2, p. 30-33, jan. 2001<br />

Santos, Milton. The Brazilian urbanization. São Paulo: Hucitec, 1993. 157p.<br />

Scheffer, S<strong>and</strong>ra Maria. Urban Space <strong>and</strong> Housing Policy: An Analysis on the Program<br />

of Plots of Prolar - Ponta Grossa. Master Thesis in Applied Social Sciences. Ponta<br />

Grossa: UEPG, 2003.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Superson et al. 2010. The deforestation of loess upl<strong>and</strong>s of SE Pol<strong>and</strong> <strong>and</strong> its stages as documented by valley deposits<br />

425<br />

The deforestation of loess upl<strong>and</strong>s of SE Pol<strong>and</strong> <strong>and</strong> its stages as<br />

documented by valley deposits (case study: Bystra river valley,<br />

Lublin Upl<strong>and</strong>)<br />

Józef Superson, Jan Reder <strong>and</strong> Wojciech Zgłobicki<br />

Institute of Earth Sciences, Maria-Curie Sklodowska University, Lublin, Pol<strong>and</strong><br />

Abstract<br />

Agriculture developing in the loess upl<strong>and</strong>s of SE Pol<strong>and</strong> caused the reduction of woodl<strong>and</strong><br />

areas, which resulted in an increased intensity of erosion processes. A greater amount of mineral<br />

deposits was supplied to the bottoms of river valleys. The objective of this study is to link these<br />

deposits with stages of prehistoric settlement <strong>and</strong> deforestation in the Bystra drainage basin.<br />

Sediments stored in the valley bottom were analysed in details. Several phases of deforestation<br />

related to human activity were described. An attempt to correlate changes of type,<br />

characteristics <strong>and</strong> geochemistry of analysed sediments with stages of deforestation was made.<br />

Keywords: deforestation, loess areas, Pol<strong>and</strong>, valley deposits<br />

1. Introduction<br />

In loess areas, forests are the best form of l<strong>and</strong> cover preventing soil erosion <strong>and</strong> gully erosion<br />

(Rodzik et al. 2009). Agriculture developing in the loess upl<strong>and</strong>s of SE Pol<strong>and</strong> from as early as<br />

the Neolithic caused the reduction of woodl<strong>and</strong> areas, which resulted in an increased intensity of<br />

erosion processes. In consequence, mainly as a result of gully erosion, a greater amount of<br />

mineral deposits was supplied to the bottoms of river valleys. Favourable natural conditions, as<br />

well as the long-term <strong>and</strong> multi-stage nature of erosion processes related to deforestation <strong>and</strong><br />

agricultural use of l<strong>and</strong>, have caused the accumulation of thick series of deposits in the bottom<br />

of valleys. They are represented by terrace deposits <strong>and</strong> alluvial fan deposits covering them<br />

(Superson, Zglobicki 2005).<br />

The objective of the study is to link these deposits with the archaeologically documented stages<br />

of prehistoric settlement <strong>and</strong> deforestation in the Bystra drainage basin based on the findings of<br />

archaeological research conducted so far <strong>and</strong> the authors’ own geomorphological studies.<br />

Deposits filling the bottom of the Bystra valley may be used as a peculiar geoarchive<br />

documenting the development stages of the environment in the area under study, including the<br />

changes in the forest cover.<br />

2. Research area <strong>and</strong> methods<br />

Bystra is a small river flowing into the Vistula in the area where the latter breaks through the<br />

upl<strong>and</strong> belt of central Pol<strong>and</strong> (Figure 1). The lower reaches of the Bystra drainage basin are<br />

situated within the loess-covered meso-region, the Naleczow Plateau (Lublin Upl<strong>and</strong>). One of<br />

the Plateau’s characteristic features is the occurrence of the loess cover that is up to 30 m thick.<br />

This meso-region is an undulating plateau with the highest elevations ranging from 180 to 220<br />

m a.s.l. The Plateau is dissected by the valleys of the Bystra <strong>and</strong> its tributaries, up to 125 m a.s.l.<br />

deep, as well as numerous trough-shaped valleys <strong>and</strong> a very dense network of gullies – up to 11<br />

km·km -2 (Maruszczak 1973). The loess areas, from the Boreal period of the Holocene, were<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Superson et al. 2010. The deforestation of loess upl<strong>and</strong>s of SE Pol<strong>and</strong> <strong>and</strong> its stages as documented by valley deposits<br />

426<br />

covered by multi-species, broad-leaved climax forests consisting of elm, lime, oak, ash <strong>and</strong><br />

hazel (Ralska-Jasiewiczowa 1991). A typical community growing in gullies nowadays are Tilio-<br />

Carpinetum forests; tree st<strong>and</strong>s are composed of hornbeam (Carpinus betulus) with an<br />

admixture of small-leaved lime (Tilia cordata), Norway maple (Acer platanoides) <strong>and</strong><br />

pedunculate oak (Quercus robur). The Bystra drainage basin has been used for agriculture since<br />

the early Neolithic (Gurba 1960). Prehistoric settlement networks in this meso-region,<br />

reconstructed through archaeological research, rank among the most developed in the Lublin<br />

Upl<strong>and</strong>, in Pol<strong>and</strong> <strong>and</strong> even in Central Europe.<br />

Based on the available archaeological material, an attempt was made to determine the stages of<br />

the deforestation of the Bystra basin. The intensity of this process in the past may only be<br />

established in qualitative terms <strong>and</strong> only based on the ascertained or supposed features of<br />

agriculture in individual cultures. More detailed quantitative analyses of changes of forested<br />

areas are only possible for the last 200 years because relatively precise comparative<br />

cartographic material is available for this period. Deposits in the bottom of the Bystra valley<br />

were studied during detailed geomorphological <strong>and</strong> geological investigations (numerous<br />

drillings), on the basis of which the character <strong>and</strong> age of the deposits was determined<br />

(radiocarbon dating). The vertical grain-size distribution of mineral deposits <strong>and</strong> selected<br />

geochemical characteristics (heavy metal content) were also determined.<br />

3. Results <strong>and</strong> discussion<br />

3.1 Stages of deforestation<br />

The deforestation of the Bystra drainage basin began in the later stage of the Atlantic period of<br />

the Holocene. It was linked with the influx of the early Neolithic people of the Lublin-Volhynia<br />

culture into this area, dated at approximately 3400 BC (Kadrow <strong>and</strong> Zakoscielna 2000). Those<br />

people cultivated small areas located in the neighbourhood of settlements, i.e. raised terraces<br />

<strong>and</strong> lower-lying, flattened promontories of the loess plateau adjoining the valley. Deforestations<br />

in that period were linked with a tremendous amount of effort. Since the use of slash-<strong>and</strong>-burn<br />

methods was limited by the natural humidity of valley bottoms <strong>and</strong> slopes, mechanical treefelling<br />

methods were used. It can be assumed that the agricultural activity of the Lublin-<br />

Volhynia culture did not cause considerable changes in the l<strong>and</strong>scape of loess promontories.<br />

The deforestation affected small areas in the immediate neighbourhood of settlements.<br />

Approximately 2900 BC, the Funnelbeaker culture people arrived in the Bystra basin. The<br />

economic activity of those people encompassed the entire area of the basin. Vast forest areas<br />

were burnt out, <strong>and</strong> fallowing was used to a great extent. The extensive use of the slash-<strong>and</strong>burn<br />

agriculture resulted in an increased area of cultivated l<strong>and</strong> <strong>and</strong> dispersal of settlements.<br />

Deforestation occurred throughout the Bystra drainage basin, but it was not permanent.<br />

In contrast with the dense Neolithic settlement in the Bystra basin, the settlement process in the<br />

Bronze Age was less intensive <strong>and</strong> developed in two stages: the early Bronze Age, after 1700<br />

BC (Trzciniec culture), <strong>and</strong> the late Bronze Age, after 1300 BC (Lusatian culture). A small<br />

number of dispersed settlements developed in a few places on terraces <strong>and</strong> in the immediate<br />

neighbourhood of the valley bottom. Their economic activity was dominated by animal<br />

husb<strong>and</strong>ry (particularly in the later stage), while farming played a secondary role. Settlement<br />

sites of the Trzciniec culture usually occur in places that were not occupied by previous<br />

Neolithic settlements, which may indicate that the previously used l<strong>and</strong> became overgrown on a<br />

considerable scale (Taras 1995). Areas affected by deforestation were small, but the process<br />

may have been relatively permanent.<br />

In the Early Iron Age (from 650 BC to 400 AD), the loess areas in the Bystra basin remained<br />

beyond the reach of the regular ecumene. Settlement sites are very sparse, located exclusively<br />

along the Bystra river, which is interpreted as proof of a very limited human penetration of the<br />

drainage basin <strong>and</strong> a migratory nature of settlement (Stasiak-Cyran 2000).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Superson et al. 2010. The deforestation of loess upl<strong>and</strong>s of SE Pol<strong>and</strong> <strong>and</strong> its stages as documented by valley deposits<br />

427<br />

Early medieval settlement in the Bystra basin dates back to the 7 th century. In the pre-statehood<br />

period (until the 11 th century), numerous, sometimes large settlements were established in the<br />

bottom of the Bystra valley <strong>and</strong> its tributaries, <strong>and</strong> deforestation in these areas must have been<br />

significant. An abrupt increase in the number of settlements can be observed from the mid-11 th<br />

century. They encroached on the loess plateau top, which resulted in relatively extensive<br />

deforestations at this l<strong>and</strong>scape level (Hoczyk-Siwkowa 1991).<br />

Starting from the 14 th century, the agricultural use of l<strong>and</strong> in the Bystra basin can be observed,<br />

varying over time, but increasingly intensive. Trees were felled in order to acquire new arable<br />

l<strong>and</strong> as well as for building <strong>and</strong>, later, industrial purposes, which led to a considerable<br />

deforestation of the drainage basin as early as the 15 th <strong>and</strong> 16 th centuries. It is estimated that the<br />

deforestation rate in the Lublin region was approximately 20% in 1340 <strong>and</strong> as much as 50% in<br />

1580 (Maruszczak 1988).<br />

3.2. Information recorded in deposits<br />

Anthropogenic changes in the l<strong>and</strong> cover of the Bystra drainage basin influenced the intensity of<br />

the supply of mineral deposits to the valley bottom. Geological <strong>and</strong> geomorphological studies<br />

show that the top of the peat layer, dating back to the Atlantic period, was partially covered as<br />

early as the Neolithic by the loam deposits of alluvial fans <strong>and</strong> by alluvial soils. Neolithic<br />

alluvial fan deposits, up to 5 m thick, were found at the mouth of extensive gully systems in the<br />

Celejow area. The sedimentation of these deposits should be linked with the deforestation of<br />

part of the drainage basin, resulting from the large-area slash-<strong>and</strong>-burn farming conducted by<br />

the people of the Funnelbeaker culture. It was probably then that the first gullies, linked with<br />

paths to water sources, formed along the axis of the trough-shaped valleys (Rodzik et all. 2009).<br />

Those gullies dissected the valley loams that subsequently were deposited in the form of alluvial<br />

fans covering a peat layer.<br />

The subsequent stage of filling the valley bottom with mineral deposits was not initiated until<br />

the early Middle Ages, which is indicated by radiocarbon dating carried out for organic<br />

formations covered by the alluvial fan at the southern valley side of the Bystra. For a long time,<br />

from the end of the Neolithic to the early Middle Ages, mineral deposits probably did not reach<br />

the bottom of the river valley or reached it in small quantities. It resulted from the sparse<br />

settlement network in the Naleczow Plateau in that period.<br />

The deposits forming the alluvial fans <strong>and</strong> the floodplain of the Bystra valley in the early<br />

Middle Ages are lithologically varied <strong>and</strong> correspond to the deposits on the slopes of the valley.<br />

These facts indicate short-distance, transverse transport as well as a varied erosion dynamics of<br />

water. At that time, gullies that had formed in the Neolithic probably became deeper, while new<br />

gullies developed in the deforested areas.<br />

The considerable deforestation of the Bystra drainage basin in the 15 th <strong>and</strong> 16 th centuries led to<br />

the formation of dense net of gullies along the axis of dry <strong>and</strong> trough-shaped valleys. Due to the<br />

absence of a consistent <strong>and</strong> permanent vegetation cover on vast areas <strong>and</strong> owing to the humidity<br />

of the medieval Climate Optimum, large amounts of material, mainly loams <strong>and</strong> s<strong>and</strong>y loams,<br />

were deposited in the bottom of the Bystra valley. These deposits form the bulk of alluvial fans<br />

<strong>and</strong> the so-called anthropogenic alluvial soil. The Heldensfeld map of 1804 shows that most<br />

contemporary gully systems were already established by then, which attests to the very high<br />

intensity of gully erosion in the Middle Ages. The map also indicates that the afforestation of<br />

certain gully system basins was considerably higher than today. The cutting down of forests in<br />

those basins in the 19 th <strong>and</strong> 20 th centuries is reflected in the deposits that make up the alluvial<br />

fans, namely the top layer of loams that, in some places, overlies poorly developed fossil soils.<br />

The last millennium was a period of substantial changes in the environment of the western part<br />

of the Lublin Upl<strong>and</strong> as well as other parts of Pol<strong>and</strong> (Maruszczak 1988). Progressive<br />

deforestation has resulted in increased dynamics of geomorphological processes in slope<br />

systems. The intensity of gully erosion, that represents the main source of material supplied to<br />

river valleys, has increased noticeably (Maruszczak 1973, 1988, Schmitt et all. 2006, Zglobicki<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Superson et al. 2010. The deforestation of loess upl<strong>and</strong>s of SE Pol<strong>and</strong> <strong>and</strong> its stages as documented by valley deposits<br />

428<br />

<strong>and</strong> Rodzik 2007). <strong>Change</strong>s in the l<strong>and</strong> cover have been reflected both in the sedimentological<br />

features of deposits as well as their geochemistry (heavy metal content). Deforestation along<br />

with the accompanying denudation is an important factor causing an increase in the content of<br />

elements in deposits.<br />

The last millennium has seen the sedimentation of mineral deposits, i.e. loams, s<strong>and</strong>y loams <strong>and</strong><br />

silty s<strong>and</strong>s in the Bystra river valley bottom. The characteristic feature was the occurrence of<br />

recurrent episodes of changes in certain parameters of the deposits, which can be linked with<br />

episodes of intensive gully erosion (Figure 2). Deposits older than 1000 years were<br />

characterised by a larger mean grain-size, ranging from 2 to 4 Φ (s<strong>and</strong>), whereas in the upper<br />

parts of the profiles (younger than 1000 years), the mean size was between 4 <strong>and</strong> 7 Φ (loams).<br />

Diversity also occurred in the case of st<strong>and</strong>ard deviation. In the lower part of the profiles, the<br />

sorting was very poor, 2-3 Φ, whereas in younger deposits it was poor or very poor, 1-3 Φ.<br />

Similar patterns were observed in the case of other valleys in the western part of the Lublin<br />

Upl<strong>and</strong> (Table 1). A small decrease in grain-size <strong>and</strong> slightly better sorting may suggest a<br />

modification of the source of material supplied to the fluvial systems under study, which results<br />

from the intensive supply of gully erosion products with the predominant silt fraction to the<br />

bottoms of river valleys.<br />

It was established that heavy metal content in deposits, particularly lead <strong>and</strong> copper, has<br />

increased over the last millennium in all studied profiles. Minor changes occurred in the case of<br />

cadmium <strong>and</strong> zinc. It has to be emphasized that the geochemical background was distinctly<br />

exceeded only in the case of surface samples (0-10 cm), which should be linked with the<br />

geochemical contamination of the environment caused by the development of industry in the<br />

area (the last 50 to 80 years). Heavy metal enrichment rates in the last millennium were as<br />

follows (the ratio between concentration at the top <strong>and</strong> concentration at the bottom of a layer<br />

dating back to the last 1000 years): Cd 1.9-2.3; Cu 1.4-2.6; Pb 1.4-3.1 <strong>and</strong> Zn 0.7-2.4 (Zglobicki<br />

2008). The concentration of the heavy metals in deposits aged more than 1000 years<br />

corresponded quite well with values characteristic of the geochemical background. The mean<br />

value of changes in geochemical characteristics, although distinct in vertical profiles, was not<br />

significant in terms of absolute values. An analysis of curves representing the trends of changes<br />

in heavy metal concentration in river deposits in the western part of the Lublin Upl<strong>and</strong> over the<br />

last 1000 years shows that compared to the 10 th century, the contemporary values of heavy<br />

metal content are between 2.5 (Pb) <strong>and</strong> 15 (Cd) times higher. In the case of lead <strong>and</strong> cadmium,<br />

the geochemical background value was exceeded around the 10 th -11 th century, which can be<br />

directly linked to the reduction of the forest cover in the area under study (Figure 3).<br />

References<br />

Gurba J. 1960. Neolithic settlements on the Lublin Loess Upl<strong>and</strong>. Annales UMCS, B, 15: 211-<br />

233.<br />

Hoczyk-Siwkowa S. 1991. Malopolska polnocno-wschodnia w VI-X wieku. Struktury osadnicze.<br />

UMCS, Lublin<br />

Kadrow S. <strong>and</strong> Zakoscielna A. 2000. An outline of the evolution of Danubian Cultures in<br />

Małopolska <strong>and</strong> Western Ukraine. Baltic-Pontic Studies, 9: 187-255.<br />

Maruszczak H. 1973. Erozja wawozowa we wschodniej czesci pasa wyzyn poludniowopolskich.<br />

Zesz. Probl. Post. Nauk Rol., 151: 15-30.<br />

Maruszczak H. 1988. Zmiany środowiska przyrodniczego kraju w czasach historycznych. In:<br />

Leszek Starkel (Ed.) Przemiany środowiska geograficznego Polski. Ossolineum, 109-135.<br />

Ralska-Jasiewiczowa M. 1991. Ewolucja szaty roslinnej. In: Leszek Starkel (Ed.) Geografia<br />

Polski.. Środowisko przyrodnicze. PWN, Warszawa, 106-127.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Superson et al. 2010. The deforestation of loess upl<strong>and</strong>s of SE Pol<strong>and</strong> <strong>and</strong> its stages as documented by valley deposits<br />

429<br />

Rodzik J., Furtak T. <strong>and</strong> Zglobicki W. 2009. The impact of snowmelt <strong>and</strong> heavy rainfall runoff<br />

on erosion rates in a gully system, Lublin Upl<strong>and</strong>, Pol<strong>and</strong>. Earth Surf. Procesess <strong>and</strong><br />

L<strong>and</strong>forms, 34: 1938–1950.<br />

Schmitt A., Rodzik J., Zgłobicki W., Russok Ch., Dotterweich M. <strong>and</strong> Bork H.-R. 2006. Time<br />

<strong>and</strong> scale of gully erosion in the Jedliczny Dol gully system, south-east Pol<strong>and</strong>. Catena,<br />

68: 24-132.<br />

Stasiak-Cyran M. 2000. Pod wpływem dwoch wielkich kultur przełomu er: celtyckiej i<br />

rzymskiej. In: Ewa Banasiewicz-Szykuła (Ed.) Archeologiczne odkrycia na terenie<br />

Kazimierskiego Parku Krajobrazowego. Lubelski Wojewodzki Konserwator Zabytkow,<br />

Lublin: 33-40.<br />

Superson J. <strong>and</strong> Zglobicki W. 2005. Rozwoj holocenskich stozkow naplywowych w dolinie<br />

Bystrej (Plaskowyz Naleczowski). In: Adam Kotarba, Kazimierz Krzemień <strong>and</strong> Jolanta<br />

Święchowicz. (Eds.) Wspołczesne ewolucja rzezby Polski. VII Zjazd Geomorfologów<br />

Polskich, UJ, Kraków, 423-429.<br />

Taras H. 1995. Kultura trzciniecka w międzyrzeczu Wisly, Bugu i Sanu. UMCS, Lublin, 262 pp.<br />

Zglobicki W. <strong>and</strong> Rodzik J. 2007: Heavy metals in slope deposits of loess areas of the Lublin<br />

Upl<strong>and</strong> (E Pol<strong>and</strong>). Catena, 71: 84-95.<br />

Zglobicki W. 2008. Geochemiczny zapis dzialalnosci człowieka w osadach stokowych i<br />

rzecznych. UMCS, Lublin, 240 p.<br />

Table 1: The diversity of mean grain-size distribution <strong>and</strong> geochemical parameters of alluvia in the<br />

western part of the Lublin Upl<strong>and</strong> (Zgłobicki 2008)<br />

Parameter Contemporary alluvia Alluvia aged up to<br />

1000 years<br />

Alluvia older than<br />

1000 years<br />

Mean grain-size [Φ] 5.2 5.9 4.5<br />

St<strong>and</strong>ard deviation [Φ] 1.9 1.7 2.3<br />

Cd content [mg/kg] 0.8 0.3 0.4<br />

Cu content [mg/kg] 14.7 10.7 6.8<br />

Pb content [mg/kg] 32.4 16.0 20.0<br />

Zn content [mg/kg] 42.9 34.7 34.0<br />

Figure 1: Location of studied area<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Superson et al. 2010. The deforestation of loess upl<strong>and</strong>s of SE Pol<strong>and</strong> <strong>and</strong> its stages as documented by valley deposits<br />

430<br />

Figure 1: Vertical distribution of heavy metals content <strong>and</strong> selected indexes of grain size distribution in<br />

profile Rablow 1 (Zglobicki 2008)<br />

Figure 2: <strong>Change</strong>s in heavy metal content in sediments of western part of Lublin Upl<strong>and</strong> <strong>and</strong> main stages<br />

of deforestation (grey arrows) during last 10 000 years (Zglobicki 2008)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.L. Teixido et al. 2010. Impacts of changes in l<strong>and</strong> use <strong>and</strong> fragmentation patterns on Atlantic coastal forests<br />

431<br />

Impacts of changes in l<strong>and</strong> use <strong>and</strong> fragmentation patterns on Atlantic<br />

coastal forests in northern Spain<br />

Alberto L. Teixido * , Luis G. Quintanilla, Francisco Carreño & David Gutiérrez<br />

1 Área de Biodiversidad y Conservación, Departamento de Biología y Geología,<br />

Universidad Rey Juan Carlos, Tulipán s/n, Móstoles, E-28933 Madrid<br />

Abstract<br />

In managed northern coast of Spain, negative consequences of forested l<strong>and</strong>scape changes have<br />

not been quantified. In one coastal forest we evaluated the l<strong>and</strong>scape transitions <strong>and</strong><br />

fragmentation patterns over time (1957-2003) by digitalising orthoimages with a high spatial<br />

resolution. Major changes on l<strong>and</strong> cover were transitions to eucalypt plantations, which showed<br />

the largest increase in area (197%). <strong>Forest</strong> declined 20% <strong>and</strong> represented 30% of l<strong>and</strong>scape area<br />

in 2003. <strong>Forest</strong> decline was mostly due to eucalypt plantations <strong>and</strong> to a water reservoir, although<br />

there were some gains by shrubl<strong>and</strong> <strong>and</strong> cultural fields been recolonised. <strong>Forest</strong> patches<br />

declined in size <strong>and</strong> core area, <strong>and</strong> increased in edge length, mean distance <strong>and</strong> in adjacency,<br />

mainly to eucalypts. The results suggested an increase in the degree of forest fragmentation <strong>and</strong><br />

changes in the matrix surrounding forest patches. This study also shows that l<strong>and</strong> use changes,<br />

mostly from eucalypt plantation intensification, negatively affected forested habitats.<br />

Keywords: Eucalypt plantations, <strong>Forest</strong> patches, L<strong>and</strong> cover classes, Riparian forest,<br />

Transitions<br />

1. Introduction<br />

Fragmentation <strong>and</strong> forest loss are one of the most threats for biodiversity (Fahrig 2003). Both<br />

processes involve patch size decrease, edge length increase <strong>and</strong> isolation, modifying the<br />

viability of populations (Hanski 1999). Despite of several studies have quantified these changes<br />

on different types of forests (e.g. Echeverria et al. 2006; Gaveau et al. 2007), in Spain montane<br />

<strong>and</strong> Mediterranean forests have been only considered (e.g. García et al. 2005), whereas Atlantic<br />

coastal forests have suffered a large fragmentation not quantified. These forests have been<br />

historically altered due to traditional l<strong>and</strong> uses <strong>and</strong> eucalypt plantations (Eucalyptus spp.)<br />

(Saura <strong>and</strong> Carballal 2004). Fragas do Eume Natural Park (NW Spain) is one of the biggest<br />

Atlantic coastal forests in Europe <strong>and</strong> has suffered a large fragmentation. In addition, a water<br />

reservoir building has altered the riparian forests, which contain several threatened species. The<br />

objectives of this study are: (1) to contribute to the underst<strong>and</strong>ing of the patterns of forest loss<br />

<strong>and</strong> fragmentation in the coastal l<strong>and</strong>scapes in northern Spain, (2) to quantify changes <strong>and</strong><br />

transitions that occurred among l<strong>and</strong> cover classes between 1957 <strong>and</strong> 2003 to underst<strong>and</strong> how<br />

they have affected to space-temporal configuration of forest <strong>and</strong>, (3) to characterise forest<br />

fragmentation patterns using st<strong>and</strong>ard l<strong>and</strong>scape metrics <strong>and</strong> adjacency with other l<strong>and</strong> cover<br />

classes as well as to determine the temporal change of those patterns.<br />

* Corresponding author. Tel.: +34 91 488 8290 - Fax:+34 91 664 7490<br />

Email address: alberto.teixido@urjc.es<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.L. Teixido et al. 2010. Impacts of changes in l<strong>and</strong> use <strong>and</strong> fragmentation patterns on Atlantic coastal forests<br />

432<br />

2. Methodology<br />

2.1 Study area<br />

Fragas do Eume NP covers 8 962 ha <strong>and</strong> it is located in the Galician dorsal range. The reserve<br />

spans 43º20’-43º26’N, <strong>and</strong> 7º52’-8º08’W. There are three native forest types depending on<br />

geomorphology: (1) oak forests, located on slopes; (2) alder forests, located on riversides with<br />

flooding sediments; <strong>and</strong> (3) hazel forests, located on steep riversides with rocky river-beds.<br />

Alder <strong>and</strong> hazel forests represent the riparian forests, <strong>and</strong> they contain threatened species such<br />

as the vertebrates Chioglossa lusitanica <strong>and</strong> Galemys pyrenaicus, <strong>and</strong> the ferns Culcita<br />

macrocarpa, Trichomanes speciosum <strong>and</strong> Woodwardia radicans.<br />

2.2 Aerial image processing <strong>and</strong> l<strong>and</strong> cover classes<br />

We used American Army’s aerial photographs from flights for the years 1956-1957 (hereafter<br />

1957) <strong>and</strong> digital orthoimages of the Geographic Information System of Agrarian Plots for the<br />

years 2002-2003 (hereafter 2003). The 1957 aerial photographs consisted of 28 contact copies<br />

(24×24 cm) with a scale of ca. 1:30 000. The photographs were scanned at ca. 2.55 m resolution,<br />

<strong>and</strong> they were subsequently georeferenced with the software ERDAS IMAGINE 8.7 using 30-<br />

40 ground control points for each frame. Orthoimages from 2003 had a spatial resolution of 0.25<br />

m. The imagery set was implemented into a GIS based on ArcGIS 9.1.<br />

Two maps were drawn showing the different l<strong>and</strong> cover classes in the Fragas do Eume NP in<br />

1957 <strong>and</strong> 2003. L<strong>and</strong> cover classification followed the system of the CORINE L<strong>and</strong> Cover 5th<br />

level project (IGN 2002). Table 1 shows the classes we digitised. Riparian forests consisted of<br />

extremely thin patches that it was practically impossible to digitise them as polygons. However,<br />

given the importance of that habitat for the persistence of key red-listed <strong>and</strong> protected species,<br />

we conducted a specific analysis for riparian forests by digitising the intersection between<br />

forests <strong>and</strong> river courses as polylines to generate a map of riparian forest length.<br />

2.3 <strong>Forest</strong> fragmentation analysis<br />

Quantification <strong>and</strong> temporal comparison of the spatial configuration of forest patches were<br />

conducted based on a set of st<strong>and</strong>ard l<strong>and</strong>scape metrics reported in recent forest fragmentation<br />

studies (e.g. Echeverria et al. 2006): (1) largest patch index (%), (2) mean patch size (ha), (3)<br />

total edge length (km), (4) total core area (ha), (5) largest patch core area (ha), (6) mean distance<br />

(m) <strong>and</strong> (7) adjacency index (total -km- <strong>and</strong> relative -%-). To calculate core area, the interior<br />

forest was defined at a distance to edge of 50 m. In the case of riparian forest, quantification <strong>and</strong><br />

temporal comparison was based on lineal segments rather on areas. For each of the two dates,<br />

we calculated the number of segments <strong>and</strong> the absolute <strong>and</strong> relative length of riparian forest<br />

from the polyline maps. We quantified all absolute <strong>and</strong> relative changes in l<strong>and</strong> cover areas <strong>and</strong><br />

lengths by calculating the difference between the values in 2003 <strong>and</strong> 1957.<br />

3. Results<br />

14 classes in 1957 <strong>and</strong> 12 in 2003 were identified (Table 1, Fig. 1). <strong>Forest</strong> suffered a 20%<br />

decrease in area <strong>and</strong> more than a two-fold increase in the number of patches between over time.<br />

In contrast, eucalypt plantations exhibited a 200% increase in area, accompanied by a moderate<br />

increase in the number of patches. Table 1 shows the overall l<strong>and</strong> cover patterns <strong>and</strong> changes.<br />

In 1957, 86% of total forest area concentrated on a large patch of 2 848 ha located along the<br />

Eume river gorge (Fig. 1); the remaining forest area occurred in patches that were smaller than<br />

1000 ha, with only ca. 4% in patches under 100 ha. The 2 848 ha patch from 1957 suffered 28%<br />

area loss as well as fragmentation into three patches of 273, 751 <strong>and</strong> 1 003 ha. The 1 003 ha<br />

patch was the largest one in 2003 <strong>and</strong> was located along the low Eume River gorge (Fig. 1). The<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.L. Teixido et al. 2010. Impacts of changes in l<strong>and</strong> use <strong>and</strong> fragmentation patterns on Atlantic coastal forests<br />

433<br />

Table 1: Number of patches, absolute (ha) <strong>and</strong> relative (%) areas for each l<strong>and</strong> cover class in Fragas do<br />

Eume Natural Park in 1957 <strong>and</strong> 2003. Net change (absolute <strong>and</strong> relative) in number of patches <strong>and</strong> area is<br />

also shown. <strong>Change</strong> values greater than 0 indicate gains, <strong>and</strong> those less than 0 indicate losses. * indicates<br />

that relative change was not calculated because that l<strong>and</strong> cover class was absent in 1957<br />

L<strong>and</strong> cover class (abbreviation) 1957 2003 <strong>Change</strong> 1957-2003<br />

Patches Area Patches Area Patches Area<br />

(ha) (%) (ha) (%) Absolute<br />

Relative<br />

(%)<br />

<strong>Forest</strong> (<strong>Forest</strong>) 64 3,306.5 36.9 137 2,659.5 29.7 +73 -647.0 -19.6<br />

Eucalypt plantations (Eucal) 162 615.3 6.9 200 1,829.7 20.4 +38 +1,214.4 +197.4<br />

Pine plantations 0 0.0 0.0 96 144.9 1.6 +96 +144.9 *<br />

Transitional woodl<strong>and</strong> (Trans) 0 0.0 0.0 267 302.9 3.4 +267 +302.9 *<br />

Gorse <strong>and</strong> heathl<strong>and</strong> (Gorse) 169 3,483.0 38.9 213 2.386.2 26.6 +44 -1,096.8 -31.5<br />

Bare rocks (Bare) 356 254.6 2.8 255 94.2 1.1 -101 -160.4 -63.0<br />

Water courses 1 59.5 0.7 1 24.5 0.3 0 -35.0 -58.8<br />

Water reservoir (Water) 0 0.0 0.0 1 400.1 4.4 +1 +400.1 *<br />

Meadows <strong>and</strong> pastures (Mead) 23 121.2 1.4 30 268.0 3.0 +7 +156.8 +129.4<br />

Complex cultivation patterns<br />

(Comp) 89 1,080.0 12.0 139 799.8 8.9 +50 -280.2 -25.9<br />

Discontinuous urban fabric 69 17.7 0.2 72 26.0 0.3 +3 +8.3 +46.9<br />

Roads 51 11.1 0.1 60 17.9 0.2 +9 +6.8 +61.3<br />

Construction sites 1 10.8 0.1 0 0.0 0.0 -1 -10.8 -100.0<br />

Mineral extraction sites 1 1.6 0.0 4 7.4 0.1 +3 +5.8 +362.5<br />

Industrial or commercial units 1 0.5 0.0 1 0.7 0.0 0 +0.2 +40.0<br />

Total 928 8,961.8 100.0 1,476 8,961.8 100.0 +489 0.0 0.0<br />

Figure 1: Spatial configuration of l<strong>and</strong> cover classes in Fragas do Eume Natural Park in 1957 <strong>and</strong> 2003<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.L. Teixido et al. 2010. Impacts of changes in l<strong>and</strong> use <strong>and</strong> fragmentation patterns on Atlantic coastal forests<br />

434<br />

Table 2: L<strong>and</strong>scape metrics of forest in Fragas do Eume Natural Park in 1957 <strong>and</strong> 2003. The absolute <strong>and</strong><br />

relative change of each l<strong>and</strong>scape metric is also shown. <strong>Change</strong> values greater than 0 indicate gains, <strong>and</strong><br />

those less than 0 indicate losses<br />

L<strong>and</strong>scape metric 1957 2003 <strong>Change</strong> 1957-2003<br />

Absolute Relative (%)<br />

Largest patch index (%) 32 11 -22 -34.4<br />

Mean patch size ± SD (ha) 52 ± 344 19 ± 110 -33 -63.5<br />

Total edge length (km) 367 484 +118 +32.1<br />

Total core area (ha) 2,054 1,178 -876 -42.6<br />

Largest patch core area (ha) 1,842 530 -1312 -71.2<br />

Mean distance ± SD (m) 489 ± 315 641 ± 546 +152 +31.1<br />

number of patches under 10 ha, <strong>and</strong> particularly those under 1 ha, increased from 1957 to 2003<br />

<strong>and</strong> represented a higher proportion of the total number of patches.<br />

There was an increase in total edge length <strong>and</strong> a decline in total core area <strong>and</strong> also a decline in<br />

largest patch core area. Mean distance among patch edges increased 31%. <strong>Forest</strong> patches also<br />

showed an increase of 32% in the total adjacency index between 1957 <strong>and</strong> 2003 (Table 2).<br />

In 1957, a large proportion (89%) of total water course length was covered by riparian forests,<br />

including the whole length of the Eume River (Fig. 2). In 2003, riparian forest length decreased<br />

34%. Losses were associated to eucalypt plantations (11%), <strong>and</strong> mostly to the building of the<br />

water reservoir (23%).<br />

Figure 2: <strong>Change</strong>s in the spatial distribution of riparian forest in Fragas do Eume Natural Park<br />

4. Discussion<br />

Our results showed a high spatial heterogeneity in l<strong>and</strong> cover over the whole area in both study<br />

periods, with most l<strong>and</strong> cover classes consisting of a large number of patches interspersed with<br />

other different classes (Fig. 1). L<strong>and</strong> cover area comparisons <strong>and</strong> detailed transition analyses<br />

showed that the l<strong>and</strong>scape was highly dynamic over the 50-year period, mostly due to<br />

intensification of exotic species plantations (especially eucalypt), farming ab<strong>and</strong>onment <strong>and</strong> the<br />

building of a water reservoir. Eucalypt plantations were mostly introduced in northern Spain<br />

during the second half of the 20th century because of their high pulp productivity. Farming<br />

ab<strong>and</strong>onment also represented large l<strong>and</strong>scape changes due to depopulation of these rural<br />

lowl<strong>and</strong>s during the second half of the 20th century (Calvo-Iglesias et al. 2006).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.L. Teixido et al. 2010. Impacts of changes in l<strong>and</strong> use <strong>and</strong> fragmentation patterns on Atlantic coastal forests<br />

435<br />

The forest loss was associated with eucalypt plantations <strong>and</strong>, in lesser degree, with the building<br />

of a water reservoir. This contrasts with the high forest losses reported over the last decades,<br />

which were mostly due to logging <strong>and</strong> agriculture, particularly for tropical areas (e.g. Gaveau et<br />

al. 2007). Those factors were also important for European forests, but in historical times prior to<br />

the period of our study (Santos et al. 2002). <strong>Forest</strong> cover was similar to those for other<br />

fragmented temperate forests (Fuller 2001; García et al. 2005; but see Santos et al. 2002). Our<br />

results also showed changes in forest spatial patterns, thus suggesting an increase in the degree<br />

of forest fragmentation over time <strong>and</strong> they were qualitatively similar to those reported for other<br />

forest studies (Fuller 2001; Echeverria et al. 2006). At this point is worth noting the importance<br />

of the fine resolution (0.01 ha) used to map patches in this study More than 20% of patches<br />

were smaller than 0.1 ha <strong>and</strong> more than 60% smaller than 1 ha in any of the study years.<br />

Temporal changes in forest patterns also implied changes in the nature of the adjacent l<strong>and</strong><br />

cover classes. The more relevant changes were the increase in adjacency with eucalypts <strong>and</strong> the<br />

water reservoir. It is known that exotic eucalypts easily spread around <strong>and</strong> into native forest<br />

patches due to their high growth rate <strong>and</strong> ability to alter forest floor quality (Fabião et al. 2002).<br />

<strong>Change</strong>s in forest patterns over time <strong>and</strong> increase in forest edge length adjacent to eucalypt<br />

plantations may have negative ecological implications on forest specialists. Eucalypt plantations<br />

generate ecological disturbances, such as increases in soil dryness <strong>and</strong> erosion, which may result<br />

in declines in plant species diversity (Fabião et al. 2002) <strong>and</strong> density of some animals as the<br />

northern-Iberian endemic amphibian C. lusitanica (Vences 1993). However, a recent genetic<br />

study in the area showed low among-population genetic variation for the protected <strong>and</strong> forest<br />

specialist ferns Culcita macrocarpa <strong>and</strong> Woodwardia radicans (Quintanilla et al. 2007).<br />

Riparian forests appeared to be relatively well-conserved, because 76% of total river course<br />

length was still covered by riparian forest in 2003. However, it was one of the l<strong>and</strong> cover classes<br />

that suffered a larger decline over the study period (34%). Thus, evaluation of conservation<br />

status for habitats based only on a snapshot of the l<strong>and</strong>scape could lead to equivocal conclusions.<br />

Riparian forests have a range of ecological properties which makes them particularly important<br />

components of l<strong>and</strong>scapes. Firstly, riparian forests have been acknowledged as dispersal<br />

corridors, hence facilitating individual exchange between natural habitat patches (Naiman et al.<br />

1993). Secondly, riparian forests may act as buffers for some forest specialists in fragmented<br />

l<strong>and</strong>scapes, as for instance mitigating the impacts derived from forest harvesting <strong>and</strong> exotic<br />

plantations (Homyack <strong>and</strong> Haas 2009).<br />

References<br />

Calvo-Iglesias, M.S., Crecente-Maseda, R. <strong>and</strong> Fra-Paleo, U., 2006. Exploring farmer’s<br />

knowledge as a source of information on past <strong>and</strong> present cultural l<strong>and</strong>scapes: A case<br />

study from NW Spain. L<strong>and</strong>scape Urban Planning, 78: 334-343.<br />

Echeverria, C., Coomes, D., Salas, J., Rey-Benayas, J.M., Lara, A. <strong>and</strong> Newton, A., 2006. Rapid<br />

deforestation <strong>and</strong> fragmentation of Chilean Temperate <strong>Forest</strong>s. Biological Conservation,<br />

130: 481-494.<br />

Fabião, A., Martins, M.C., Cerveira, C., Santos, C., Lousã, M., Madeira, M. <strong>and</strong> Correia, A.,<br />

2002. Influence of soil <strong>and</strong> organic residue management on biomass <strong>and</strong> biodiversity of<br />

understory vegetation in a Eucalyptus globulus Labill. plantation. <strong>Forest</strong> Ecology <strong>and</strong><br />

Management, 2002: 87-100.<br />

Fahrig, L., 2003. Effects of habitat fragmentation on biodiversity. Annual Review of Ecology<br />

Evolution <strong>and</strong> Systematics, 34: 487-515.<br />

Fuller, D., 2001. <strong>Forest</strong> fragmentation in Loudoun County, Virginia, USA evaluated with<br />

multitemporal L<strong>and</strong>sat imagery. L<strong>and</strong>scape Ecology, 16: 627-642.<br />

García, D., Quevedo, M., Ramón-Obeso, J. <strong>and</strong> Abajo, A., 2005. Fragmentation patterns <strong>and</strong><br />

protection of montane forest in the Cantabrian range (NW Spain). <strong>Forest</strong> Ecology <strong>and</strong><br />

Management, 208: 29-43.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.L. Teixido et al. 2010. Impacts of changes in l<strong>and</strong> use <strong>and</strong> fragmentation patterns on Atlantic coastal forests<br />

436<br />

Gaveau, D.L.A., W<strong>and</strong>ono, H. <strong>and</strong> Setiabudi, F., 2007. Three decades of deforestation in<br />

southwest Sumatra: Have protected areas halted forest loss <strong>and</strong> logging, <strong>and</strong> promoted regrowth<br />

Biological Conservation, 131: 495-504.<br />

Hanski, I., 1999. Metapopulation ecology. Oxford University Press, Oxford, UK.<br />

Homyack, J.A. <strong>and</strong> Haas, C.A., 2009. Long-term effects of experimental forest harvesting on<br />

abundance <strong>and</strong> reproductive demography of terrestrial salam<strong>and</strong>ers. Biological<br />

Conservation, 142: 110-121.<br />

IGN, 2002. CORINE 2000. Descripción de la nomenclatura del CORINE L<strong>and</strong> Cover al nivel<br />

5º (Diciembre 2002). CORINE L<strong>and</strong> Cover. Actualización 2000. I<strong>and</strong>CLC2000. Instituto<br />

Geográfico Nacional, Centro Nacional de Información Geográfica, Ministerio de Fomento,<br />

Madrid, Spain.<br />

Naiman, R.J., Decamps, H. <strong>and</strong> Pollock, M., 1993. The role of riparian corridors in maintaining<br />

regional biodiversity. Ecological Applications, 3: 209-212.<br />

Quintanilla, L.G., Pajarón, S., Pangua, E. <strong>and</strong> Amigo, J., 2007. Allozyme variation in the<br />

sympatric ferns Culcita macrocarpa <strong>and</strong> Woodwardia radicans at the northern extreme of<br />

their ranges. Plant Systematics <strong>and</strong> Evolution, 263: 135-144.<br />

Santos, T., Tellería, J.L. <strong>and</strong> Carbonell, R., 2002. Bird conservation in fragmented<br />

Mediterranean forests of Spain: effects of geographical location, habitat <strong>and</strong> l<strong>and</strong>scape<br />

degradation. Biological Conservation, 105: 113-125.<br />

Saura, S. <strong>and</strong> Carballal, P., 2004. Discrimination of native <strong>and</strong> exotic forest patterns through<br />

shape irregularity indices: an analysis in the l<strong>and</strong>scapes of Galicia, Spain. L<strong>and</strong>scape<br />

Ecology, 19: 647-662.<br />

Vences, M., 1993. Habitat choice of the salam<strong>and</strong>er Chioglossa lusitanica: the effects of<br />

eucalypt plantations. Amphibia-Reptilia, 14: 201-212.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Assessing multi-temporal l<strong>and</strong> cover changes in the Mata Nacional da Peneda Geres<br />

437<br />

Assessing multi-temporal l<strong>and</strong> cover changes in the Mata Nacional da<br />

Peneda Geres National Park (1995 <strong>and</strong> 2009), Portugal - a l<strong>and</strong> change<br />

modeler approach for l<strong>and</strong>scape spatial patterns modelling <strong>and</strong><br />

structural evaluation<br />

Helder Viana 1,3 & José Aranha 2,3*<br />

1<br />

Centro de Estudos em Educação, Tecnologias e Saúde, Agrarian <strong>Superior</strong> School,<br />

Polytechnic Institute of Viseu, Quinta da Alagoa, 3500-606 Viseu, Portugal<br />

2<br />

Departamento de Ciências Florestais e Arquitectura Paisagista, Universidade de Trásos-Montes<br />

e Alto Douro, 5001-801 Vila Real, Portugal<br />

3<br />

CITAB, Universidade de Trás-os-Montes e Alto Douro, 5001-801 Vila Real, Portugal<br />

Abstract<br />

The present study sought to evaluate l<strong>and</strong> cover evolution between 1995 <strong>and</strong> 2009, within the<br />

Mata Nacional of Peneda Geres (Portugal). This study was based on L<strong>and</strong>sat TM images<br />

classification <strong>and</strong> GIS procedures, such as L<strong>and</strong> <strong>Change</strong> Modeller approach. L<strong>and</strong>scape<br />

diversity <strong>and</strong> structural changes were analysed by means of Mean Shape Index, Shannon’s<br />

Diversity Index <strong>and</strong> Patch analysis, in order compare l<strong>and</strong>scape metrics <strong>and</strong> to calculate l<strong>and</strong><br />

cover dynamics. The achieved results enable to state that l<strong>and</strong> cover classes presented<br />

significant structural changes. The most significant changes occur in the l<strong>and</strong> cover classes of<br />

Pinus pinaster; Quercus robur; Acacia dealbata; <strong>and</strong> shrub l<strong>and</strong>. The most worrying result was<br />

achieved for the Acacia dealbata, which presented a strong invasive behaviour. L<strong>and</strong>scape<br />

metric analysis showed a significant stratification increasing <strong>and</strong> a dramatic reduction of patch<br />

surfaces. In spite of spatial changes observed, the achieved biodiversity indexes are very alike<br />

for both dates.<br />

Keywords: L<strong>and</strong>scape ecology, Spatial patterns analysis, <strong>Change</strong>s prediction, L<strong>and</strong>sat, Gerês<br />

1. Introduction<br />

Natural or National Parks use to be created as way to preserve wild areas <strong>and</strong> to give a chance to<br />

nature undergo its trend. However, due to centuries of anthropogenic action, nature must be<br />

under human surveillance in order to redress previous misuse or to guide ecosystem recovery.<br />

Over the last 50 years, natural areas have suffered significant changes, which are visible from<br />

l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> use changes. These changes are due to several causes: human exodus from<br />

rural areas, which leaded to uncontrolled shrub growing <strong>and</strong>, this way, to wild fire starting <strong>and</strong><br />

spread, alien species introduction <strong>and</strong> spread, climatic changes <strong>and</strong> unbalanced forestry<br />

composition.<br />

The Peneda Geres National Park is a good <strong>and</strong> living example of previous state. From the last<br />

200 years, rural population used wild l<strong>and</strong> for grazing cattle, to cut shrubs for cattle’s beds, to<br />

grow timber <strong>and</strong> to cut wood for fireplaces, during rigorous inter. However, from the last 30<br />

years, human action within the Park decreased drastically, both by changes in Portuguese way<br />

of life <strong>and</strong> restrictive policy imposed by law. This actions combination is now very visible in<br />

* Corresponding author. Tel.: + 00 351 259 350 856 - Fax: + 00 351 250 350 480<br />

Email address: j.aranha.utad@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Assessing multi-temporal l<strong>and</strong> cover changes in the Mata Nacional da Peneda Geres<br />

438<br />

l<strong>and</strong>scape <strong>and</strong> leaded to changes in ecosystem balance. Due to a strong anthropogenic pass, this<br />

area must be under continuous surveillance in order to avoid irreversible ecosystem losses <strong>and</strong><br />

unbalanced evolution.<br />

This is also visible throughout Europe, at different scales <strong>and</strong> domains, due to changes in<br />

technologies, policy <strong>and</strong> economical growing up (Verburg et al. 2008).<br />

L<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> use dynamic analysis <strong>and</strong> evolution qualification has to be used as a way to<br />

estimate environmental consequences due to l<strong>and</strong>scape changes (López et al. 2001, Flamenco-<br />

S<strong>and</strong>ovala, et al. 2007; Rutherford et al., 2008). Multi temporal spatial pattern analysis <strong>and</strong><br />

quantitative l<strong>and</strong>scape structural analysis have been successful used to derive research in this<br />

domain (O´Neill et al. 1988; Bresee et al. 2001, Tischendorf 2001).<br />

This research was derived using Remote Sensing <strong>and</strong> Geographic Information Systems methods,<br />

in order to assess the l<strong>and</strong>scape dynamics in Gerês National <strong>Forest</strong>. We evaluated the global<br />

l<strong>and</strong>scape composition <strong>and</strong> structure from 1995 to 2009 <strong>and</strong> made projections for 2015 based on<br />

transition observed from the past to 2009.<br />

Was selected Geres National Park, as it is a protected area of high l<strong>and</strong>scape value <strong>and</strong><br />

ecological, part of the Peneda-Geres (PNPG). The Geres National Park contains one of the most<br />

important oak woods, consisting predominantly of a centuries-old oak forest (carvalhal da<br />

Albergaria) where is noticed the presence of fauna <strong>and</strong> flora species characteristic of gerensiana<br />

formation. PNPG, it is classified as a Zone of Partial Protection of Natural Environment Area<br />

<strong>and</strong> Rede Natura 2000 (Anonymous, 1995). In general, this area can be classified as an area of<br />

mountainous altitude, since about 82% of the territory, has altitudes ranging between 700 <strong>and</strong><br />

1400 m.<br />

Due to its morphological characteristics, which result from the conjunction of four mountains<br />

(Peneda, Soajo, Amarela e Gerês), this area creates a natural barrier to hot <strong>and</strong> wet air coming<br />

from Atlantic Ocean. This results in high values of mean annual rainfall, ranging from 2400 mm<br />

to 2800 mm, or 3000 mm in very wet years. This is the wettest area in Portugal <strong>and</strong> even in<br />

Europe (Fontes, 2005). In this territory we can find several forest species, with high<br />

productivities as pines, eucalyptus, oaks <strong>and</strong> much other conifers <strong>and</strong> deciduous species. The<br />

most species are well adapted <strong>and</strong> in equilibrium in the ecosystem, however, many introduced<br />

species became invaders (e.g. Acacia dealbata, Acacia melanoxilon).<br />

2. Methodology<br />

Figure 1: Study area location<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Assessing multi-temporal l<strong>and</strong> cover changes in the Mata Nacional da Peneda Geres<br />

439<br />

They were used satellite images (L<strong>and</strong>sat-5 TM, 5/July/1995, L<strong>and</strong>sat-5 ETM+, 1/April/2001,<br />

L<strong>and</strong>sat-5 TM, 4/August/2006 <strong>and</strong> L<strong>and</strong>sat-5 ETM+, 16/June/2009), ancillary data such as:<br />

topographic maps at the scale of 1:25 000 with a 10m contour interval, orthophotomaps from<br />

1995, 2000 <strong>and</strong> 2005 at a scale of 1:10 000, cartographic elements in vector format such as<br />

roads, rivers, administrative boundaries <strong>and</strong> environmental characteristics.<br />

In previous intensive fieldwork, information about l<strong>and</strong> cover classes <strong>and</strong> ground control points<br />

(GCP) were collected using a DGPS (Differential GPS). The GCP were collected in road<br />

crosses, barrages or other notable points perfectly visible in the images (Toutin 2004). This data<br />

was used as auxiliary tools in the geometric correction, in the definition of training classes <strong>and</strong><br />

in the validation stage (Lilles<strong>and</strong> et al. 2004, Toutin 2004, Eastman 2006, Scally 2006, D’ Iorio<br />

et al. 2007, Tsai 2007<br />

A digital elevation model (DEM) was created from 10m interval contours, collected from the<br />

1:25000 topographic maps of representing the study area, in order to calculate slope <strong>and</strong> aspect<br />

<strong>and</strong> to apply images topographic normalization.<br />

The conceptual framework of the research included pre-processing stage, image enhancement,<br />

image transformation (RGB composition, vegetation indices calculation <strong>and</strong> principal<br />

component analysis), image classification <strong>and</strong> interpretation <strong>and</strong> accuracy assessment. It was<br />

used IDRISI 32 (Eastman 2006) image processing software for image data processing, <strong>and</strong><br />

ArcGis 9.x (ESRI 2004) software for GIS based analyses procedures.<br />

During previous field work for data collection, it was used a DGPS (Differential <strong>Global</strong><br />

Positioning System) in training areas mapping (e.g. coniferous st<strong>and</strong>s, burnt areas, acacia areas,<br />

etc.) <strong>and</strong> ground control points collection (e.g. roads intersection). DGPS data was corrected<br />

with Trimble® GPS Pathfinder® Office software.<br />

The adopted l<strong>and</strong> use/cover scheme, used in image classification, was based upon Corine L<strong>and</strong><br />

Cover classification (CLC2000).<br />

In a first image classification stage, they were used nine l<strong>and</strong> cover classes, in order to create a<br />

general l<strong>and</strong> cover map. In a second stage, using a local scale <strong>and</strong> after fieldwork with a DGPS<br />

(Differential <strong>Global</strong> Positioning System) for accuracy assessment, l<strong>and</strong> cover maps were<br />

updated during, in order to assign the correct forestry class name to each l<strong>and</strong> cover class. In<br />

total, fifteen l<strong>and</strong> use/cover classes were considered in this study:<br />

• Shrubs,<br />

• Pinus sylvestris,<br />

• Acacia dealbata,<br />

• Quercus robur;<br />

• Mixed Broadleaves;<br />

• Mixed Coniferous;<br />

• Mixed Coniferous <strong>and</strong> Broadleaves;<br />

• Chamaecyparis lawsoniana,<br />

• Mixed Shrubs <strong>and</strong> Quercus sp.;<br />

• Fagus sylvatica;<br />

• Acacia melanoxylon;<br />

• Pinus pinaster;<br />

• Pinus nigra;<br />

• Arbutus unedo; <strong>and</strong><br />

• Betula celtiberica.<br />

They were performed several automatic classification methods, including unsupervised models<br />

<strong>and</strong> Principal Component Analysis (Asner 1998, Song et al 2001, D’ Iorio et al. 2007, Tsai et al.<br />

2007). Supervised classification of multispectral images was performed, running the Maximum<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Assessing multi-temporal l<strong>and</strong> cover changes in the Mata Nacional da Peneda Geres<br />

440<br />

Likelihood classifier (MLC) <strong>and</strong> the Minimum Distance to Means classifier (MDMC)<br />

(Lilles<strong>and</strong> et al. 2004, Eastman 2006, Scally 2006).<br />

The accuracy of a classified image refers to the extent to which it agrees with a set of reference<br />

data. Thus, an error matrix was created in order to compare the accuracy of maps obtained from<br />

satellite images classification. The error matrix provides a mean to calculate the overall<br />

accuracy <strong>and</strong> to compute accuracies of each category (Congalton <strong>and</strong> Green 1999).<br />

It was calculated Kappa statistic (Cohen, 1960), because of its ability to provide information<br />

about a single matrix <strong>and</strong> to statistically compare matrices, in order to get another measure of<br />

agreement between the predicted values <strong>and</strong> the observed values, the, (Cohen, 1960, Rosenfield<br />

<strong>and</strong> Fitzpatrick-Lins 1986, Congalton <strong>and</strong> Green 1999, Meidinger, 2003).<br />

For l<strong>and</strong> cover changes detection, it was used a pixel-to-pixel comparison of classified images,<br />

because it is a method widely used <strong>and</strong> easily understood. This step preformed by L<strong>and</strong> <strong>Change</strong><br />

Modeler (LCM - IDRISI Andes, Eastman, 2006) <strong>and</strong> aimed to compare the images generated<br />

for the different years of study.<br />

L<strong>and</strong> <strong>Change</strong> Modeller (LCM - IDRISI Andes, Eastman, 2006) enable l<strong>and</strong>scape changes<br />

analysis; Shaping the potential transition of the l<strong>and</strong> cover classes, provide the direction of<br />

changes in the future, assessment to their implications for biodiversity <strong>and</strong> to evaluate plans of<br />

action for ecological sustainability maintenance.<br />

At the final stage, l<strong>and</strong> cover maps, created from 1995 to 2009 satellite images, were submitted<br />

to pattern analysis in order to assess l<strong>and</strong>scape structural quantification. This stage was<br />

performed by means of Patch Analyst 4 (Rempel, 2008), which is a working modulo available<br />

in ArcGIS 9.x GIS software.<br />

3. Result<br />

The observed changes, in terms of net percentage change, are more significant in classes: mixed<br />

conifer, Shrubs with oaks self seeding, Arbutus unedo, Betula Celtiberica, Acacia dealbata,<br />

melanoxylon Acacia, Quercus robur.<br />

The most important modification was recorded in oak areas (Quercus robur) with an effective<br />

reduction of about 444 ha, representing a decrease of 113.9%. Classes of increasing l<strong>and</strong> cover<br />

are: Acacia dealbata increased more than 200ha, <strong>and</strong> Acacia melanoxylon, with more than 30ha.<br />

The substitution of pure Quercus robur st<strong>and</strong>s by shrubs, in a such large area, was due to strong<br />

shrub’s developed, which is replacing the younger oak areas. It was also noticed that some<br />

Quercus robur some pure st<strong>and</strong>s, classified in 1995, were classified in 2009 as deciduous <strong>and</strong><br />

coniferous mixed st<strong>and</strong>s, which can lead to gradual replacement of deciduous trees on these<br />

areas.<br />

Analysing the global balance of Acacia dealbata changes (gains <strong>and</strong> losses), Acacia dealbata<br />

increased its area over Pinus pinaster pure st<strong>and</strong>s (80ha) <strong>and</strong> shrub areas (120ha).<br />

The Acacia melanoxylon increasing area was made, as for Acacia dealbata, by the replacement<br />

of Pinus pinaster pure st<strong>and</strong>s (30ha), <strong>and</strong> these changes correspond to about 8% decrease of<br />

pine in the global balance of gains <strong>and</strong> losses in forest occupation of this class<br />

Despite the total area assigned to Pinus pinaster st<strong>and</strong>s (polygons <strong>and</strong> 35 319ha in 1995 to 64<br />

292ha in 2009 polygons) <strong>and</strong> to shrub l<strong>and</strong> (3580ha polygons <strong>and</strong> 88 in 1995; 3232ha polygons<br />

<strong>and</strong> 522 in 2009) had not varied greatly, in absolute values, the spatial distribution has suffered<br />

a large increasing in fragmentation. In the case of shrub l<strong>and</strong>, it was observed that some areas<br />

have be replaced by a mix of young oaks <strong>and</strong> shrubs. The Pinus pinaster st<strong>and</strong>s were replaced<br />

by scattered trees <strong>and</strong> shrubs, as well by Acacia dealtata, as previous presented. In Table 1 are<br />

presented the calculated results for metrics of diversity <strong>and</strong> inter dispersion for 1995 <strong>and</strong> 2009.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Assessing multi-temporal l<strong>and</strong> cover changes in the Mata Nacional da Peneda Geres<br />

441<br />

4. Discussion<br />

Table 2: Metrics of diversity <strong>and</strong> inter dispersion for 1995 <strong>and</strong> 2009<br />

Metrics Acronym 1995 2009<br />

Mean shape index MSI 1.56 1.60<br />

Area weighted mean shape index AWMSI 5.22 6.42<br />

Mean polygon fractal dimension MPFD 1.07 1.08<br />

L<strong>and</strong>scape shape index LSI 8.46 10.42<br />

Metrics of diversity <strong>and</strong> inter dispersion<br />

Mean distance to nearest neighbour (m) MNN 282.50 186.50<br />

Mean proximity index MPI 771.39 1680.10<br />

Index of dispersion <strong>and</strong> overlapping (%) IDO 43.61 59.20<br />

Shannon diversity index SDI 1.02 1.13<br />

Simpson diversity index SIDI 0.47 0.46<br />

Shannon equity diversity index SEI 0.38 0.42<br />

Simpson equity diversity index SIEI 0.50 0.49<br />

Modified Simpson diversity index MSIDI 0.63 0.61<br />

Modified Simpson equity diversity index MSIEI 0.23 0.23<br />

Results from this research showed that the vegetal l<strong>and</strong> cover within Mata Nacional of Peneda<br />

Geres National Park is under high changing dynamic. For the period in analysis, some l<strong>and</strong><br />

cover classes evidenced significant lost (e.g. Quercus robur) <strong>and</strong> significant gains (e.g. Acacia<br />

dealbata).<br />

The achieved results for Quercus robur, the ex-libres of Mata Nacional of Peneda Geres<br />

National Park showed a strong area decrease, with tendency to be smaller in a new future. This<br />

is a strong cut in l<strong>and</strong>scape value <strong>and</strong> biodiversity.<br />

Open areas within Quercus robur st<strong>and</strong>s mapped in 1995 were now occupied by Pinus pinaster,<br />

Fagus sylvatica, shrubs <strong>and</strong> Acacia dealbata.<br />

In general, Pinus pinaster <strong>and</strong> shrubs presented a stabilised l<strong>and</strong> cover area, these classes<br />

evidenced some dynamic, but the total areas assigned to both classes are near the same in both<br />

dates.<br />

One of the most significant results was achieved for alien trees (Acacia dealbata <strong>and</strong> Acacia<br />

melanoxylon). According to prediction maps, these species are those with higher potential to<br />

enlarging their actual area. These tow trees species are well known because of their strong<br />

invasive behaviour, which leads to additional worry, as they can quickly colonize almost all<br />

Mata Nacional, leading to a severe ecosystem disturbance.<br />

References<br />

Anonymous (1971). Decreto nº 187/71 de 08 de Maio.<br />

Anonymous (1995). RCM nº 134/95, de 11 de Novembro.<br />

Bresee, M.K., Moine, J.L., Mather, S., Brosofske, K.D., Chen, J., Crow T.R. And Rademacher,<br />

J., 2001, Disturbance <strong>and</strong> l<strong>and</strong>scape dynamics in the Chequamegon National <strong>Forest</strong><br />

Wisconsin, USA, from 1972 to 2001. L<strong>and</strong>scape Ecology 19: 291–309.<br />

Cohen. I., 1960, A coefficient of agreement of nominal scales. Educational <strong>and</strong> Psychological<br />

Measurement, 20(1): 37-46.<br />

Congalton, R.G. And Green, K., 1999, Assessing the Accuracy of Remotely Sensed Data:<br />

Principles <strong>and</strong> practices (Boca Raton, London, New York: Lewis Publishers).<br />

Eastman, J. R., 2006, Idrisi Andes. Guide to GIS <strong>and</strong> Image Processing. Clark University. USA.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Assessing multi-temporal l<strong>and</strong> cover changes in the Mata Nacional da Peneda Geres<br />

442<br />

Flamenco-S<strong>and</strong>ovala, A., Martínez R.M. And Raúl M.O., 2007, Assessing implications of l<strong>and</strong>use<br />

<strong>and</strong> l<strong>and</strong>-cover change dynamics for conservation of a highly diverse tropical rain<br />

forest. Biological conservation 138, 131 – 145.<br />

Fontes, A., 2005, Modelação do Risco de Incêndio no Parque Nacional da Peneda Gerês;<br />

Seminário de Investigação em Geografia Física e Ambiente; UM/PNPG; Guimarães.<br />

ICN, 1995. Plano de Ordenamento do Parque Nacional da Peneda-Gerês. Relatório de<br />

Síntese. Parque Nacional Da Peneda-Gerês. Instituto da Conservação da Natureza. Braga.<br />

Lilles<strong>and</strong>, T.M., Kiefer, R.W. And Chipman, J.W., 2004, Remote Sensing <strong>and</strong> image<br />

Interpretation. Fifth Edition, John Wiley <strong>and</strong> Sons Inc., New York.<br />

López, E., Bocco, G., Mendoza, M., And Duhau, E., 2001, Predicting l<strong>and</strong>-cover <strong>and</strong> l<strong>and</strong>-use<br />

change in the urban fringe. A case in Morelia city, Mexico. L<strong>and</strong>scape <strong>and</strong> Urban<br />

Planning 55, 271- 285.<br />

Luijten, J.C., 2003, A systematic method for generating l<strong>and</strong> use patterns using stochastic rules<br />

<strong>and</strong> basic l<strong>and</strong>scape characteristics: results for a Colombian hillside watershed.<br />

Agriculture, Ecosystems <strong>and</strong> Environment 95, 427–441.<br />

O´Neill, R., Krummel, J., Gardner, R., Sugihara, G., Jackson, B., Deangelis, D., Milne, B.,<br />

Turner, M., Zygmunt, B., Christensen, S., Dale, V. And Graham, R., 1988, Indices of<br />

l<strong>and</strong>scape pattern. L<strong>and</strong>scape Ecology, 1, 153-162.<br />

Rempel, R., 2008, Centre for Northern <strong>Forest</strong> Ecosystem Research (Ontário Ministry of Natural<br />

Resources), Lakehead University Campus, Thunder Bay, Ontario. [on-line] Available<br />

online at: http://flash.lakeheadu.ca/~rrempel/patch/index.html. (Acessed 13.04.2009)<br />

Rosenfield, G. H. & Fitzpatrick-Lins, K., 1986, A coefficient of agreement as a measure of<br />

thematic accuracy. Photogrammetric Engineering <strong>and</strong> Remote Sensing, 52(2), pp. 223-<br />

227.<br />

Rutherford, G. N., Bebi, P., Edwards, P.J. And Zimmermann, N.E., 2008, Assessing l<strong>and</strong>-use<br />

statistics to model l<strong>and</strong> cover change a mountainous l<strong>and</strong>scape in the European Alps.<br />

Ecological modelling, 212, 460–471.<br />

Scally, R., 2006, GIS for Environmental Management. ESRI Press, 380 New York Street,<br />

Redl<strong>and</strong>s, California, USA.<br />

Song, C., Woodcock, C. E., Seto, K. C., Lenney, M. P. And Macomber, S. A., 2001,<br />

Classification <strong>and</strong> <strong>Change</strong> Detection Using L<strong>and</strong>sat TM Data: When <strong>and</strong> How to Correct<br />

Atmospheric Effects Remote Sensing of Environment, 75, pp. 230-244.<br />

Tischendorf, L., 2001, Can l<strong>and</strong>scape indices predict ecological processes consistently<br />

L<strong>and</strong>scape Ecology 16: 235–254.<br />

Toutin, T., 2004, Review article: Geometric processing of remote sensing images: models,<br />

algorithms <strong>and</strong> methods. International Journal of Remote Sensing, 25(10), 1893-1924.<br />

Tsai, F., Lin, E. K. And Yoshino, K., 2007, Spectrally segmented principal component analysis<br />

of hyperspectral imagery for mapping invasive plant species. International Journal of<br />

Remote Sensing, 28, 1023-1039.<br />

Verburg, P.H., Eickhout,•B. And Meijl, H.V., 2008, A multi-scale, multi-model approach for<br />

analyzing the future dynamics of European l<strong>and</strong> use. Special Issue Paper. Annals of<br />

regional science 42:57–77.<br />

Acknowledgement<br />

Authors woul like to expresse is acknowledge to Fundação para a Ciência e Tecnologia (FCT),<br />

project REEQ/1163/AGR/2005, CITAB - UTAD <strong>and</strong> programme SFRH/PROTEC/49626/2009<br />

who support this work.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Mapping invasive species (Acacia dealbata Link)<br />

443<br />

Mapping invasive species (Acacia dealbata Link) using<br />

ASTER/TERRA <strong>and</strong> LANDSAT 7 ETM+ imagery<br />

Helder Viana 1,3 & José Aranha 2,3*<br />

1<br />

Centro de Estudos em Educação, Tecnologias e Saúde, Agrarian <strong>Superior</strong> School,<br />

Polytechnic Institute of Viseu, Quinta da Alagoa, 3500-606 Viseu, Portugal<br />

2<br />

Departamento de Ciências Florestais e Arquitectura Paisagista, Universidade de Trásos-Montes<br />

e Alto Douro, 5001-801 Vila Real, Portugal<br />

3<br />

CITAB, Universidade de Trás-os-Montes e Alto Douro, 5001-801 Vila Real, Portugal<br />

Abstract<br />

The rapid spread of invasive alien species (IAS) is now recognised as one of the greatest threats<br />

to the ecological <strong>and</strong> economic well being of the planet. This study shows a comparison<br />

between ASTER/TERRA <strong>and</strong> ETM+/LANDSAT 7 sensors data suitability for mapping the<br />

Acacia dealbata Link spots. The work was carried out in central Portugal (Viseu region) where<br />

the presence of invader species in pure st<strong>and</strong>s is quite significant. The images were orthorectified<br />

<strong>and</strong> submitted to supervised classifications techniques. The achieved results showed an<br />

overall accuracy of 89.42% over the ETM+ image <strong>and</strong> 86.69% over the ASTER image. For the<br />

class Acacia dealbata Link, the producer’s precision was 100% for both images but the user’s<br />

accuracy was only 23% in ETM+ <strong>and</strong> 12% in ASTER image. The obtained results suggest good<br />

perspectives for the use of this type of satellite images in order to detect <strong>and</strong> map this invasive<br />

species.<br />

Keywords: Alien species, Acacia dealbata, l<strong>and</strong> cover classification, L<strong>and</strong>sat ETM+, ASTER<br />

1. Introduction<br />

The rapid spread of invasive alien species (IAS) is causing irreparable damage to global<br />

ecosystems. Variously referred to as exotic, non-native, alien, noxious, or non-indigenous<br />

weeds, these species are causing enormous damage to biodiversity <strong>and</strong> to the valuable natural<br />

agricultural systems, which we depend on (Coimbra 1999; Liberal & Esteves 1999; Aguiar et al.<br />

2001; Aguiar & Ferreira 2005, Viana 2005). In Portugal, the establishment <strong>and</strong> spread of<br />

invasive species, particularly Acacia dealbata Link, has increased over time. They were<br />

introduced deliberately as silvicultural, for soil fixing, as ornamental or by another pretext,<br />

being now a serious problem for the ecosystems, with difficult control <strong>and</strong> even impossible<br />

eradication. Identifying those areas is essential to quantify the real dimension of the problem<br />

(Coimbra 1999; Liberal & Esteves 1999; Bargeron et al. 2003; Viana 2005).<br />

With the coming in sight of new image sensors, with different characteristics, <strong>and</strong> data<br />

availability, it is important to test the potentialities for specific uses as IAS detection <strong>and</strong><br />

mapping (Asner 1998; Bargeron et al. 2003; Leitão et al. 2003; Brundu 2005; Chikhaoui et al.<br />

2005; Viana 2005; D’Iorio et al. 2007).<br />

The Advanced Space borne Thermal Emission <strong>and</strong> Reflection Radiometer (ASTER) is a<br />

research facility launched on NASA’s Earth Observing System, on board of TERRA satellite<br />

(previously called EOS AM-l), in December 1999. As expected ASTER data has been used in<br />

specific areas of scientific investigation, including vegetation <strong>and</strong> ecosystem dynamics, hazard<br />

* Corresponding author; Telf. + 00 351 259 350 856 - Fax. + 00 351 250 350 480<br />

Email address: j.aranha.utad@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Mapping invasive species (Acacia dealbata Link)<br />

444<br />

monitoring, geology <strong>and</strong> soils, l<strong>and</strong> surface climatology, hydrology, <strong>and</strong> l<strong>and</strong> cover change<br />

(Abrams 2000; NASA 2004; Tangestani 2004; Chikhaoui et al. 2005; Euroimage 2008).<br />

The L<strong>and</strong>sat programme constitutes the longest data register of the L<strong>and</strong> surface from the Space.<br />

The Enhanced Thematic Mapper-Plus (ETM+) was launched on April 15, 1999 on board of<br />

L<strong>and</strong>sat7 <strong>and</strong>, as TM sensor data, imagery have been extensively used for agricultural<br />

evaluation, forest management inventories, geological surveys, water resource estimates, coastal<br />

zone appraisals, <strong>and</strong> a host of other applications (Song et al. 2001; Darvishsefat 2003;<br />

Thenkabail et al. 2004; Peterson 2005; Viana 2005; NASA 2006; NASA 2007;).<br />

Given the characteristics of ASTER sensor systems, which provide imagery data at higher<br />

spatial resolution (15m on VNIR) than ETM+ (30m), the same temporal resolution-16 days, <strong>and</strong><br />

with a unique combination of wide spectral coverage, in this study we tested <strong>and</strong> compared both<br />

imagery performance in the mapping of a specific class of forest l<strong>and</strong> cover (Acacia dealbata<br />

Link).<br />

Study Area was a 64Km x 60Km rectangle in the region of Viseu (centre of Portugal) (see<br />

Figure 1). It’s a heterogeneous area with a complex topography <strong>and</strong> fragmented l<strong>and</strong> cover, with<br />

elevation in the range of 100 to 1800m; high climatic variability, with annual mean precipitation<br />

in the range of 800 to 2800 mm <strong>and</strong> annual mean temperatures of < 7.5 to 16 ºC.<br />

Figure 1: Study area location.<br />

2. Methodology<br />

2.1. Data acquisition<br />

The study was developed using multispectral images covering Viseu’s region, in Portugal,<br />

provided from the sensor ETM+/L<strong>and</strong>sat 7, <strong>and</strong> sensor ASTER/Terra (L1b format), on the<br />

VNIR b<strong>and</strong>s. The acquisition date of ETM+ was on 24, January 2003, period in which these<br />

plants were flowery <strong>and</strong> ASTER on 7, October 2003, since it was the available image closest to<br />

the ETM+ acquisition date. Topographic maps 1:25000 <strong>and</strong> orthophotomap 1:10000 were used<br />

as auxiliary tools in the definition of training classes <strong>and</strong> in the validation stage. The collection<br />

of spatial information as cartographic elements e.g. l<strong>and</strong> cover classes, roads <strong>and</strong> ground control<br />

points (GCP) was done by GPS. A total of 85 plots of Acacia dealbata were measured in a sum<br />

of 66.6 hectares, with a mean area around 0.78 hectares, later used for training classes. The GCP<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Mapping invasive species (Acacia dealbata Link)<br />

445<br />

were collected in road crosses, barrages or other notable points visible in the images (Lilles<strong>and</strong><br />

et al 2004; Viana 2005, Eastman 2006).<br />

2.2. Data processing<br />

The conceptual framework of the research followed 5 central steps: geometric correction, Image<br />

enhancement, image transformation (vegetation indices <strong>and</strong> principal component analysis),<br />

classification <strong>and</strong> interpretation <strong>and</strong> validation. DGPS (Differential <strong>Global</strong> Positioning System)<br />

data was corrected with Pathfinder Office, image data were processed with IDRISI 32, <strong>and</strong> GIS<br />

based analyses was done with ArcGis software.<br />

In first place the ASTER data (level 1B) of VNIR b<strong>and</strong>s (1, 2, 3N), with 15m spatial resolution,<br />

in the HDF format, <strong>and</strong> ETM+ data of pan b<strong>and</strong> with 15 m <strong>and</strong> multispectral b<strong>and</strong> (1~5, 7) with<br />

30 m spatial resolution were imported to IDRISI. The ASTER image <strong>and</strong> pan ETM+ images<br />

were registered with GCP, <strong>and</strong> the multispectral ETM+ b<strong>and</strong>s were based on image-to-image<br />

method, using the already registered images as reference.<br />

For image classification, it was adopted a l<strong>and</strong> use/cover scheme based upon the Corine L<strong>and</strong><br />

Cover classification (CLC2000). They were performed automatic classification methods,<br />

unsupervised models <strong>and</strong> Principal Component Analysis (Song et al. 2001, Tsai et al. 2007).<br />

Supervised classification of multispectral images was performed, running the Maximum<br />

Likelihood classifier (MLC) <strong>and</strong> the Minimum Distance to Means Classifier (MDMC)<br />

(Lilles<strong>and</strong> et al. 2004, Eastman 2006, Scally 2006). The accuracy of a classified image refers to<br />

the extent to which it agrees with a set of reference data. Thus, an error matrix was created in<br />

order to compare the accuracy of maps obtained from satellite images classification. The error<br />

matrix provides a mean to calculate the overall accuracy <strong>and</strong> to compute accuracies of each<br />

category (Congalton <strong>and</strong> Green 1999). Kappa statistic (Cohen 1960), because of its ability to<br />

provide information about a single matrix <strong>and</strong> to statistically compare matrices, was calculated<br />

in order to get another measure of agreement between the predicted values <strong>and</strong> the observed<br />

values, the, (Cohen 1960, Rosenfield <strong>and</strong> Fitzpatrick-Lins 1986, Congalton <strong>and</strong> Green 1999,<br />

Meidinger, 2003).<br />

For l<strong>and</strong> cover changes detection, it was used a pixel-to-pixel comparison of classified images,<br />

because it is a method widely used <strong>and</strong> easily understood.<br />

3. Result<br />

After supervised image classification, the resulting images area very alike. These results were<br />

evaluated using a set of 2304 validation points <strong>and</strong> error matrix. The overall statistics of<br />

classifications are summarised in the Tables 1 <strong>and</strong> 2.<br />

Table 1: Producer’s <strong>and</strong> User’s for ETM+ <strong>and</strong> ASTER imagery classification<br />

Producer’s accuracy (%) User’s accuracy (%)<br />

L<strong>and</strong> cover class<br />

ETM+ ASTER ETM+ ASTER<br />

<strong>Forest</strong>ed areas 93.1 94.7 99.9 97.8<br />

Meadow 81.8 56.0 96.8 84.8<br />

Acacia dealbata 100.0 100.0 22.4 11.1<br />

As previous presented table show, both images provided quite similar results. The best<br />

classification was achieved with the Maximum Likelihood classifier (Table 2). Although the<br />

ETM+ image achieve a higher overall accuracy, <strong>and</strong> superior user’s accuracy, for all the<br />

considered l<strong>and</strong> cover classes, only the Md class had higher producer accuracy (81.8% <strong>and</strong><br />

56.0%, respectively).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Mapping invasive species (Acacia dealbata Link)<br />

446<br />

Table 2: Overall accuracy of ETM+ <strong>and</strong> ASTER imagery classification<br />

Method Overall accuracy (%) Kappa statistics<br />

ETM+ - MLC 89.42 0.8543<br />

ETM+ - MDMC 63.53 0.5190<br />

ASTER - MLC 86.69 0.8121<br />

ASTER - MDMC 69.21 0.5688<br />

For the l<strong>and</strong> cover class Fr the reliability of ETM+ image (93.1%) was minor than ASTER<br />

(94.7%). In the best classification (MLC), the class “acacias” had shown a Producer’s accuracy<br />

of 100% in both ETM+ <strong>and</strong> ASTER images. This happened due to the commission error being<br />

77.57% in ETM+ <strong>and</strong> 88.89% in ASTER image (Table 3). This means that the vector shapes<br />

considered in the creation of this spectral signature were representative of the Ac class, <strong>and</strong> had<br />

been well created, however given the nature of this l<strong>and</strong> cover class (permanent leaf <strong>and</strong> closed<br />

canopy) some pixels belonged to other l<strong>and</strong> cover class were classified as Ad, principally Md<br />

class in reason of their similar spectral response.<br />

All Ad (Acacia dealbata) spots mapped with a DGPS were well classified by satellite image<br />

classification. However, do to small spot dimension <strong>and</strong> fragmented l<strong>and</strong>scape, some Md<br />

(Meadow) were misclassified as Ad (Acacia dealbata) areas.<br />

4. Discussion<br />

In this paper/work we have compared ASTER <strong>and</strong> ETM+ data in forest applications. The<br />

accuracy of image classification <strong>and</strong> interpretation was tested <strong>and</strong> compared. The resulting<br />

conclusions are:<br />

- ASTER data can be registered with elevated accuracy with error less than half pixel.<br />

- ASTER is better than ETM+ data in visual surface feature identification.<br />

- ASTER classification has the same effect as ETM+ with high accuracy;<br />

- With ASTER it was possible to classify l<strong>and</strong> cover shapes with smaller areas in reason of their<br />

superior spatial resolution.<br />

- A superior resolution in ASTER (15m) is not an evident advantage when mapping features<br />

with reduced dimension such as Ad (Acacia dealbata), given that the spectral confusion, fact<br />

amplified in fractionated l<strong>and</strong>scapes as in the Centre of Portugal.<br />

- The Maximum Likelihood classifier gave better results than the Minimum Distance to Means<br />

classifier in the supervised classification, involving l<strong>and</strong> cover classes (acacias) distributed in<br />

parcels with small areas.<br />

- Given the uncertainty about follow-on L<strong>and</strong>sat ETM+ sensor, ASTER imagery could be<br />

supply suitable images for monitoring applications, with similar results.<br />

References<br />

Abrams, M., 2000, The Advanced Spaceborne Thermal Emission & Reflection Radiometer<br />

(ASTER): Data products for the high spatial resolution imager on NASA's Terra platform.<br />

International Journal of Remote Sensing, 21(5), pp. 847-859.<br />

Aguiar, F. C., Ferreira, M. T. & Moreira, I., 2001, Exotic & Native Vegetation Establishment<br />

Following Channelization of a Western Iberian River. Regulated Rivers: Research &<br />

Management, 17, pp. 509-526.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Mapping invasive species (Acacia dealbata Link)<br />

447<br />

Almeida, T. I. R., De Souza Filho, C. R. & Rossetto, R., 2006, ASTER & L&sat ETM+ images<br />

applied to sugarcane yield forecast. International Journal of Remote Sensing, 27(19),<br />

pp.4057-4069.<br />

Atlas Do Ambiente, 1978, Carta da Distribuição de Acácias e Eucaliptos. Escala 1/1000000.<br />

In: Atlas do Ambiente Digital, Instituto do Ambiente. Impressa no Instituto Hidrográfico,<br />

Lisboa.<br />

Bargeron, C. T., Moorhead, D. J., Douce, G. K., Reardon, R. C. & Miller, A. E. (Tech.<br />

Coordinators), 2003, Invasive Plants of the Eastern U.S.: Identification & Control. [CD-<br />

ROM]. USDA <strong>Forest</strong> Service - <strong>Forest</strong> Health Technology Enterprise Team. Morgantown,<br />

WV USA. FHTET.<br />

Brundu, G., Camarda, I. & Satta, V., 2003, A methodological approach for mapping alien plants<br />

in Sardinia (Italy). In Plant Invasions: Ecological Threats & Management Solutions,<br />

Child, L.E.; Brock, J.H.; Brundu, G.; Prach, K.; Pysek, P.; Wade, P.M.; Williamson, M.<br />

(Eds.), pp. 41-62. (Leiden, The Netherl&s: Backhuys Publishers).<br />

Chavez, P. S., Jr. (1996), Image-based atmospheric corrections revisited & improved.<br />

Photogrammetric Engineering & Remote Sensing. 62, pp. 1025–1036.<br />

Chavez, P. S., Jr., 1988, An improved dark-object subtraction technique for atmospheric<br />

scattering correction of multi- spectral data. Remote Sensing of Environment, 24, pp. 459–<br />

479.<br />

Chavez, P. S., Jr., 1989, Radiometric calibration of L&sat Thematic Mapper multispectral<br />

images. Photogrammetric Engineering & Remote Sensing. 55, pp.1285–1294.<br />

Chikhaoui, M.; Bonn, F., Bokoye, A. I. & Merzouk, A., 2005, A spectral index for l&<br />

degradation mapping using ASTER data: Application to a semi-arid Mediterranean<br />

catchment. International Journal of Applied Earth Observation & Geoinformation, 7,<br />

pp.140-153.<br />

Civco, D. L., 1989, Topographic normalization of L&sat Thematic Mapper digital imagery.<br />

Photogrammetric Engineering & Remote Sensing, 55, pp. 1303–1309.<br />

Cohen. I., 1960, A coefficient of agreement of nominal scales. Educational & Psychological<br />

Measurement, 20(1): 37-46.<br />

Coimbra, A. J. M., 1999, Distribuição da mimosa (Acacia dealbata Link) na área do Parque<br />

Natural da Serra da Estrela. In 1º Encontro sobre Invasoras Lenhosas, 16 a 18 de<br />

Novembro, Parque Nacional da Peneda-Gerês, Gerês, pp. 186-189.<br />

CONGALTON, R.G. & K. Green., 1999, Assessing the Accuracy of Remotely Sensed Data:<br />

Principles & Practices. Lewis Publishers. 137 pp.<br />

Coutinho, A. X. P., 1939, Flora de Portugal (Plantas Vasculares). 2ª Edição dirigida por R. T.<br />

Palhinha. (Lisboa, Bertr& Irmãos Ltd.).<br />

Du, Y., Teillet, P. M., & Cihlar, J., 2002, Radiometric normalization of multitemporal highresolution<br />

satellite images with quality control for l& cover change detection. Remote<br />

Sensing of Environment, 82, pp. 123–134.<br />

Eastman, J. R., 2006, Idrisi Andes. Guide to GIS & Image Processing. (Worcester: Clark Labs<br />

for Cartographic Technology & Geographic Analysis, Clark University).<br />

Esri, 2004, ArcGIS version 9.2 (Redl&s, California: ESRI).<br />

Fensham, R. J., Fairfax, R. J., Holman, J. E. & P. J. Whitehead, 2002, Quantitative assessment<br />

of vegetation structural attributes from aerial photography. International Journal of<br />

Remote Sensing, 23(11), pp. 2293–2317.<br />

Franco, J. A., 1971, Nova Flora de Portugal (Continente e Açores). 1ª. ed, Lisboa (Impressa na<br />

Sociedade Astória Lda).<br />

Godinho-Ferreira, P., Azevedo, A. & Rego, F., 2005, Carta da Tipologia Florestal de Portugal<br />

Continental. Silva Lusitana, 13(1), pp. 1-34.<br />

Larsson, H., 2002, Acacia canopy cover changes in Rawashda forest reserve, Kassala Province,<br />

Eastern Sudan, using linear regression NDVI models. International Journal of Remote<br />

Sensing, 23(2), pp. 335 339.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


H. Viana & J. Aranha 2010. Mapping invasive species (Acacia dealbata Link)<br />

448<br />

Leitão, P., Pinto, M. & Máguas, C., 2003, Remote sensing evaluation of post-fire s<strong>and</strong>-dune<br />

acacia forest dynamics. In Invasão de Habitats Naturais por espécies lenhosas<br />

infestantes, 7 de Novembro, Parque Nacional da Peneda-Gerês, Gerês.<br />

Liberal, M. & Esteves, M., 1999, Invasão de Acacia dealbata no Parque Nacional da Peneda-<br />

Gerês. In 1º Encontro sobre Invasoras Lenhosas, 16 a 18 de Novembro, Parque Nacional<br />

da Peneda-Gerês, Gerês, pp 99-103.<br />

LILLES&, T.M., KIEFER, R.W. & CHIPMAN, J.W., 2004, Remote Sensing & Image<br />

Interpretation, 5th edition, 765 pp. (New York, NY: John Wiley & Sons).<br />

Lu, D. & Weng, Q., 2007, A survey of image classification methods & techniques for<br />

improving classification performance. International Journal of Remote Sensing Vol. 28,<br />

No. 5, pp. 823–870.<br />

Meyer, P., Itten, K.I., Kellenberger, T., S&meier, S. & S&meier, R., 1993, Radiometric<br />

corrections of topographically induced effects on L&sat TM data in an alpine<br />

environment. ISPRS Journal of Photogrammetry & Remote Sensing, 48, pp. 17–28.<br />

Montserud, R. A. & Leamans, R., 1992, Comparing global vegetation map with the Kappa<br />

statistic. Ecological Modeling, 62, pp. 275–293.<br />

Peterson, E. B., 2005, Estimating cover of an invasive grass (Bromus tectorum) using tobit<br />

regression & phenology derived from two dates of L<strong>and</strong>sat ETM+ data. International<br />

Journal of Remote Sensing, 26(12), pp. 2491–2507.<br />

Rosenfield, G. H. & Fitzpatrick-Lins, K., 1986, A coefficient of agreement as a measure of<br />

thematic accuracy. Photogrammetric Engineering & Remote Sensing, 52(2): 223-227.<br />

Song, C., Woodcock, C. E., Seto, K. C., Lenney, M. P. & Macomber, S. A., 2001, Classification<br />

& <strong>Change</strong> Detection Using L&sat TM Data: When & How to Correct Atmospheric<br />

Effects Remote Sensing of Environment, 75, pp. 230-244.<br />

Tangestani, M. H., Mazhari, N., Agar, B. & Moore, F., 2008, Evaluating Advanced Spaceborne<br />

Thermal Emission & Reflection Radiometer (ASTER) data for alteration zone<br />

enhancement in a semiarid area, northern Shahr-e-Babak, SE Iran. International Journal<br />

of Remote Sensing, 29(10), pp. 2833–2850<br />

Teillet, P.M., Guindon, B. & Goodenough, D.G., 1982, On the slope-aspect correction of<br />

multispectral scanner data. Canadian Journal of Remote Sensing, 8, pp. 84–106.<br />

Toutin, T., 2004, Review article: Geometric processing of remote sensing images: models,<br />

algorithms & methods. International Journal of Remote Sensing, 25(10), pp. 1893-1924.<br />

Tsai, F., Lin, E. K. & Yoshino, K., 2007, Spectrally segmented principal component analysis of<br />

hyperspectral imagery for mapping invasive plant species. International Journal of<br />

Remote Sensing, 28, pp. 1023-1039.<br />

USGS, 2008, ASTER Overview. NASA's Earth Observing System (EOS). L<strong>and</strong> Processes<br />

Distributed Active Archive Centre (LPDAAC). Available online at:<br />

https://lpdaac.usgs.gov/lpdaac/products/aster_overview (accessed 10 January 2009).<br />

USGS, 2009, L<strong>and</strong>sat Enhanced Thematic Mapper Plus (ETM+). Product information.<br />

Available online at: http://eros.usgs.gov/products/satellite/l&sat7.php (accessed 10<br />

January 2009).<br />

Viana, H., 2005, Aplicação de Tecnologias de Detecção Remota e Sistemas de Informação<br />

Geográfica na Identificação de Áreas Ocupadas com Acacia dealbata Link. Dissertação<br />

de Mestrado em Engenharia dos Recursos Florestais. University of Trás-os-Montes e Alto<br />

Douro.<br />

Acknowledgement<br />

Authors woul like to expresse is acknowledge to Fundação para a Ciência e Tecnologia (FCT),<br />

project REEQ/1163/AGR/2005, CITAB - UTAD <strong>and</strong> programme SFRH/PROTEC/49626/2009<br />

who support this work.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Section 6<br />

Tools of l<strong>and</strong>scape assessment <strong>and</strong> management


N. Avani et al. 2010. Investigation on mountain l<strong>and</strong>scape parameters on Juniper species growth<br />

450<br />

Investigation on mountain l<strong>and</strong>scape parameters on Juniper species<br />

growth (case study: Firozkooh region, Tehran)<br />

Nazi Avani 1* , Hamid Jalilv<strong>and</strong> 1 & Vahid Etemad 2<br />

1<br />

Maz<strong>and</strong>aran University, Iran<br />

2<br />

Faculty of Natural resources, Tehran University, Iran<br />

Abstract<br />

It is generally agreed that sustainable development <strong>and</strong> management of upl<strong>and</strong> natural resource<br />

for the welfare of local population should be the key objective of watershed management, which<br />

includes sustainable utilization <strong>and</strong> conservation of forest resource of community or watershed<br />

level as one of its important components. Juniperus polycarpus L. is one of the species that<br />

grows in the mountain areas, <strong>and</strong> has important role in mountain l<strong>and</strong>scape. Our purpose of this<br />

study is study of l<strong>and</strong>scape parameters on Juniper growth, that which parameters have the best<br />

effect on this species growth. Therefore in different slope <strong>and</strong> aspects, we measured parameters<br />

include total height, canopy height, canopy diameter <strong>and</strong> basal diameter. To order of study of<br />

grow of juniper species in different slopes <strong>and</strong> aspects in Aminabad region of Firozkooh<br />

(Tehran province), thrown inventory network to systematic r<strong>and</strong>om, then in study area, number<br />

of 40 sample plots (0.5 ha) was taken. Then parameters include total height, canopy height,<br />

canopy diameter <strong>and</strong> basal diameter were measured. Slope <strong>and</strong> aspect maps were supplied with<br />

GIS. The results show that greatest diameter of trees was in southeastern <strong>and</strong> the greatest height<br />

was in west aspect, <strong>and</strong> the most height <strong>and</strong> diameter was in average slopes. Therefore, for<br />

planting with this species, at first west, southeastern aspects <strong>and</strong> average slopes were suggested.<br />

Keywords: mountain l<strong>and</strong>scape, Juniperus polycarpus L., Firozkooh, Iran<br />

Introduction<br />

Iran is located in 26 until 39 degree that is sub-arid, arid <strong>and</strong> desert region in the world. In the<br />

mountain areas in Iran because of humid climate, shrub <strong>and</strong> tree species are grown. Among tree<br />

species, genus of Pistacia with famous species P. mutica L. has the largest area in Iran. After<br />

that genus of Juniperus with species J. excelsa <strong>and</strong> J. polycarpus cover the large area of<br />

elevations in Iran. *<br />

Genus of Juniperus in Iran has six species <strong>and</strong> sub- species that grow in mountain l<strong>and</strong>scapes.<br />

Juniperus occupied arid lime areas <strong>and</strong> we can see them in Alborz, Khorasan, Azarbaijan, Arak,<br />

Kerman, Balochestan <strong>and</strong> etc.<br />

Many studies were done in the world about effect of ecologic factors on evergreen <strong>and</strong> conifer<br />

species. Fisher & Gardner (1995), with studying on juniper forests in the north of Oman show<br />

that topography, hydrology <strong>and</strong> climate condition has large effect on Juniper species growth.<br />

Kanji (2000), with studying on forests in the north of Japan show that conifers just grow in the<br />

south, west or west southern aspects. Razzag (1986), soil depth, rainfall, <strong>and</strong> aspect are the<br />

effective factors on growth of Pinus eldarica. Johnson & Miller (2006), they study factors such<br />

* Corresponding Author. Tel: 00989128046062<br />

Email Address: avani.nazi@yahoo.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Avani et al. 2010. Investigation on mountain l<strong>and</strong>scape parameters on Juniper species growth<br />

451<br />

as topography <strong>and</strong> elevation on the growth of J. occidntalis in the USA. The results show that<br />

settlement <strong>and</strong> density of trees in upper elevations <strong>and</strong> in the north aspects has been increased.<br />

Moemeni moghadam (2002), the number of ecologic characteristics of Juniper species was<br />

studied in Shirvan (Khorasan province), <strong>and</strong> then the results show that slope effect on survival,<br />

number in hectare <strong>and</strong> form quotient <strong>and</strong> aspect effect on canopy height to total height <strong>and</strong> form<br />

quotient <strong>and</strong> biodiversity.<br />

Our purpose of this study is study of l<strong>and</strong>scape parameters on Juniper growth, that which<br />

parameters have the best effect on this species growth. Therefore in this study the effect of slope<br />

<strong>and</strong> aspect on the growth of Juniper species was studied.<br />

Material <strong>and</strong> methods<br />

In the way of Firozkooh (Tehran province) after Hom<strong>and</strong>, in the slopes <strong>and</strong> elevations of this<br />

region, the st<strong>and</strong>s of Juniper species were determined. Soil depth is low that is 15 to 30 cm, the<br />

altitude is 2600 until 2710 m, dominant species is Juniper in this region. The amount of bearing<br />

fruit trees is very much, but there are very low seedlings in the region. The study area is located<br />

in the Hableroad watershed, in 35◦ 42´ to 35◦ 44´ northern latitude <strong>and</strong> 52◦ 33´ to52◦ 35´ eastern<br />

longitude.<br />

For doing this study, at the first we distinguished the region in the topography map, then thrown<br />

the inventory network with 100*500 dimension <strong>and</strong> 40 plots with 0.5 ha area. After that height<br />

<strong>and</strong> diameter growth variables in the different slopes <strong>and</strong> aspects were measured. The data were<br />

analyzed in the SAS software <strong>and</strong> with SNK test.<br />

Results<br />

The results show that the most basal diameter was in southeastern (44.67 cm) <strong>and</strong> the least basal<br />

diameter was in northwestern (31.57 cm) (figure 1). The results of the means comparison show<br />

that aspect had the significant effect on the height of trees in 1% level of probability <strong>and</strong> the<br />

most height was in west (6.19m) <strong>and</strong> the least height was in southwestern (5.12m) (figure 2).<br />

The results of the means comparison in different slopes show that slope had significant effect on<br />

basal diameter of trees in 5% level of probability, <strong>and</strong> the most basal diameter was in 80% of<br />

slope (40.03 cm) <strong>and</strong> the least was in 20% of slope (27.84 cm) (figure 3).<br />

In 5% level of probability, slope had the significant effect on height of trees <strong>and</strong> the most height<br />

was in 55% of slope (7.26m) <strong>and</strong> the least height was in 15% of slope (5.03 m) (figure 4).<br />

Canopy height had the different significant in different aspects in 1% level <strong>and</strong> the most canopy<br />

height was in west (6 m) <strong>and</strong> the least canopy height was in southwestern (5.2 m) (figure 5). The<br />

most canopy diameter was in west (4.57 m) <strong>and</strong> the least was in west southern (3.57 m) (figure<br />

6).<br />

The results show that slope had the significant effect on the canopy diameter of trees in 5%<br />

level that the most canopy diameter was in 55% of slope (4.57 m) <strong>and</strong> the least was in 20%<br />

slope (3.6 m) (figure 7). Slope in 5% level of probability had the significant effect on canopy<br />

height <strong>and</strong> the most was in 55% of slope (7.3 m) <strong>and</strong> the least was in 15% of slope (5.04 m)<br />

(figure 8).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Avani et al. 2010. Investigation on mountain l<strong>and</strong>scape parameters on Juniper species growth<br />

452<br />

Discussion <strong>and</strong> conclusion<br />

Because these species was grown in mountain regions <strong>and</strong> in the hard condition as climatic <strong>and</strong><br />

etc., in these regions for assessment growth variables, factors such as basal diameter because of<br />

b coppice, canopy diameter <strong>and</strong> height because of much distance between trees that cause<br />

increasing diameter growth, were measured.<br />

The results of this study shows that aspect <strong>and</strong> slope as mountain l<strong>and</strong>scape parameters had the<br />

significant effect on basal diameter, total height, canopy height <strong>and</strong> canopy diameter. The<br />

highest trees were in the west <strong>and</strong> the most diameters were in southeastern aspect. Jonhson<br />

(2006), in his study on Western Juniper shows that topography is the effective factor on its<br />

growth.<br />

Kanji (2000) conifers grow in the west, south <strong>and</strong> south western. In our study too, we show that<br />

the most growth was in the west.<br />

Fisher <strong>and</strong> Razzag (1995 & 1998) topography factors <strong>and</strong> aspect have significant effect on the<br />

growth of Juniper species <strong>and</strong> Pinus eldarica. In our study too, aspect had the significant effect<br />

on the Juniper growth.<br />

Therefore we can suggest that for planting with Juniper species in the mountain area for<br />

sustainable development of these regions <strong>and</strong> better management, at the first, west aspect then<br />

eastern <strong>and</strong> in average slopes was selected.<br />

References<br />

Fisher, M. & Gardner, A.S., 1995. The status <strong>and</strong> ecology of Juniperus excelsa sub. P.<br />

polycarpus woodl<strong>and</strong> in the northern of Oman. Vegation. 779: 33-48.<br />

Johnson, D.D. & Miller, R.F., 2006. Structure <strong>and</strong> development of exp<strong>and</strong>ing western juniper<br />

woodl<strong>and</strong>s as influenced by two topographic variables. <strong>Forest</strong> ecology <strong>and</strong> management.<br />

229: 7-15.<br />

Kanji, N., 2000. Edaphic controls on mosaic structure of the mixed deciduous<br />

broadleaf/conifer forest in northern Japan. <strong>Forest</strong> ecology <strong>and</strong> management. 127: 169-<br />

179.<br />

Moemeni Moghadam, T., 2002. Investigation some ecologic <strong>and</strong> silviculture characteristics of<br />

natural habitat of Juniperus in Koopedagh (Shirvan). Tarbiat Modares University.<br />

Razzag, A.A., 1986. The influence of habitat on afforestation success in Jordan. Gottinger<br />

Beitrage zur L<strong>and</strong> und <strong>Forest</strong>wirtschaft in den Tropen und Subtropen. 25:105-109.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Avani et al. 2010. Investigation on mountain l<strong>and</strong>scape parameters on Juniper species growth<br />

453<br />

Figure 1: basal diameter of trees in different aspects<br />

Figure 2: total height of trees in different aspects<br />

Figure 3: basal diameter in different slopes<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Avani et al. 2010. Investigation on mountain l<strong>and</strong>scape parameters on Juniper species growth<br />

454<br />

Figure 4: height of trees in different slopes<br />

Figure 5: canopy height in different aspects<br />

Figure 6: canopy diameter in different aspects<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Avani et al. 2010. Investigation on mountain l<strong>and</strong>scape parameters on Juniper species growth<br />

455<br />

Figure 7: canopy diameter in different slopes<br />

Figure 8: canopy height in different slopes<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.C. Azevedo et al. 2010. Spatial dynamics of chestnut blight disease at the plot level using the Ripley’s K function<br />

456<br />

Spatial dynamics of chestnut blight disease at the plot level using the<br />

Ripley’s K function<br />

João C. Azevedo 1* , Valentim Coelho 1 , João P. Castro 1 , Diogo Spínola 2 & Eugénia<br />

Gouveia 1<br />

1<br />

CIMO, Centro de Investigação de Montanha, <strong>Escola</strong> <strong>Superior</strong> Agrária, Instituto<br />

Politécnico de Bragança, Portugal<br />

2 Universidade Federal de Viçosa, Brazil<br />

Abstract<br />

We used the Ripley’s K function to describe the spatial dynamics of chestnut blight<br />

(Cryphonectria parasitica (Murrill) Barr) in sweet chestnut orchards to look at pattern in the<br />

pathogen distribution over time <strong>and</strong> the effect of the location of infected trees on the pattern of<br />

disease spread. We used data on infected <strong>and</strong> dead trees in 2003, 2004, 2005, <strong>and</strong> 2009 in 4<br />

orchards located in Curopos parish, Portugal. We found both r<strong>and</strong>om <strong>and</strong> aggregated patterns of<br />

infected trees in the beginning of the study period <strong>and</strong> significant association of infected trees<br />

between successive dates, particularly at short distances. Two of the 4 studied orchards showed<br />

significant clustering of infected <strong>and</strong> dead trees in any of the dates observed but r<strong>and</strong>om spatial<br />

pattern in the remaining two which can possibly be explained by both natural propagation of the<br />

disease <strong>and</strong> management practices.<br />

Keywords: Cryphonectria parasitica, chestnut blight, Ripley’s K function, Portugal<br />

1. Introduction<br />

Although the recognition of the importance of the spatial dimension in the study of infectious<br />

diseases is not new, only recently significant developments in spatial pathology <strong>and</strong><br />

epidemiology took place. L<strong>and</strong>scape <strong>and</strong> spatial epidemiology emerged in the 2000’s from the<br />

application of l<strong>and</strong>scape ecological concepts <strong>and</strong> methods in the analysis of disease dispersal in<br />

animal, human, <strong>and</strong> plant hosts (Ostfeld 2005, Plantegenest et al. 2007). Similarly, l<strong>and</strong>scape<br />

pathology has grown within l<strong>and</strong>scape ecology <strong>and</strong> forest pathology dealing with large scale<br />

disease propagation processes <strong>and</strong> the ways they affect <strong>and</strong> are affected by l<strong>and</strong>scape<br />

heterogeneity (Holdenrieder et al. 2004).<br />

The spatial propagation of pathogens at small scales (e.g., forest st<strong>and</strong>) is also ecologically<br />

relevant, although seldom approached in the literature. The underst<strong>and</strong>ing of small scale<br />

epidemiology processes is of interest in explaining, modeling <strong>and</strong> forecasting pathogen related<br />

spatial processes at this particular scale as well as at larger scales such as national- or regionallevel<br />

pathogen dispersal (e.g., Kelly <strong>and</strong> Meentemeyer 2002).<br />

In this study, we analyzed spatial patterns of chestnut blight (Cryphonectria parasitica (Murrill)<br />

Barr) infected trees at the orchard level over time. The objectives were to i) investigate pathogen<br />

spread temporal <strong>and</strong> spatial pattern, <strong>and</strong> ii) analyze the effect of the location of infected trees on<br />

the spatial pattern of disease spread.<br />

* Corresponding author. Tel.: (+351) 273 303 341 - Fax: (+351) 273 325 405<br />

Email address: jazevedo@ipb.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.C. Azevedo et al. 2010. Spatial dynamics of chestnut blight disease at the plot level using the Ripley’s K function<br />

457<br />

2. Methodology<br />

We used data from 4 orchards located in the Curopos parish, Vinhais, Portugal, that have been<br />

monitored for individual tree health condition based on field surveys in 2003, 2004, 2005 <strong>and</strong><br />

2009 (Table 1; Fig 1). Common management practices in these plots included pruning, excision<br />

of cankers <strong>and</strong> replacement of dead trees.<br />

Table 1: Area of study plots <strong>and</strong> number of dead <strong>and</strong> infected trees per plot <strong>and</strong> year<br />

Plot 1 Plot 2 Plot 3 Plot 4<br />

Area (ha) 1.34 1.85 0.92 1.05<br />

Dead <strong>and</strong> infected trees (no.)<br />

2003 21 37 12 71<br />

2004 38 60 18 88<br />

2005 14 54 11 94<br />

2009 96 148 118 104<br />

We analyzed the spatial pattern of dead trees <strong>and</strong> trees presenting symptoms of blight disease<br />

with the Ripley’s K function (Ripley 1976). This second-order analysis method allows<br />

summarizing spatial patterns, fitting models to describe patterns <strong>and</strong> comparing patterns among<br />

events at variable scales (Dixon 2002). Although the method fits the point-set structure of trees<br />

at the plot level the use of Ripley’s K is very rare in plot level pathology studies.<br />

The Ripley’s K(t) is estimated as (Haase, 1995):<br />

−2<br />

−1<br />

() t = n A∑∑<br />

wij<br />

It<br />

( uij<br />

)<br />

Kˆ (1)<br />

i≠<br />

j<br />

where<br />

n is the number of individuals (locations) in the plot,<br />

A is the area of the plot (m 2 )<br />

I t is a counting variable,<br />

u ij is the distance between i <strong>and</strong> j locations, <strong>and</strong><br />

w ij is a weighting factor for edge correction purposes.<br />

The bivariate form of K(t) is estimated as (Dixon, 2002):<br />

() −<br />

n1 n2<br />

1<br />

= ∑∑<br />

− 1<br />

t n1n<br />

2<br />

A wij<br />

It<br />

( uij<br />

)<br />

K ˆ<br />

(2)<br />

i<br />

j<br />

where n 1 <strong>and</strong> n 2 are the number of individuals in the populations under comparison.<br />

K can be normalized as L(t)=√K(t)/π <strong>and</strong> represented graphically as L(t)-t as a function of t.<br />

Positive values of L(t)-t indicate aggregation of events while negative values indicate a regular<br />

pattern. In the bivariate form, positive values indicate association between populations of events<br />

<strong>and</strong> negative values indicate segregation. Zero values indicate complete spatial r<strong>and</strong>omness<br />

(Poisson) in the univariate case <strong>and</strong> no pattern in the bivariate case. Confidence intervals are<br />

created to test for significance of pattern.<br />

We used RIPPER (Feagin & Wu, personal communication) to calculate Ripley’s K with the<br />

edge correction method of Getis & Franklin (1987). Maximum distance was half side of the plot<br />

<strong>and</strong> the same box size was used in each plot for all the dates considered. 95% confidence<br />

intervals were established based upon 200 Monte Carlo simulations.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.C. Azevedo et al. 2010. Spatial dynamics of chestnut blight disease at the plot level using the Ripley’s K function<br />

458<br />

Plot 1 Plot 2<br />

Plot 3 Plot 4<br />

Figure 1. 2003, 2004, 2005, <strong>and</strong> 2009 health status of individual trees in 4 plots in the Curopos parish,<br />

Portugal.<br />

3. Resuls<br />

In Plot 1, there was statistically significant clustering of infected trees at distances above 10m in<br />

any of the surveyed years (Figure 2). The same pattern was observed in Plot 4 although the<br />

strength of clustering was much lower. Plots 2 <strong>and</strong> 3 showed weak aggregation, often nonstatistically<br />

significant. In Plot 2 there was slightly significant clustering for distances from15 to<br />

25m in 2004 <strong>and</strong> 2005 <strong>and</strong> above 10m in 2009. In Plot 3 there was significant clustering from<br />

25 to 50m in 2005 <strong>and</strong> above 10m in 2009 (Figure 3).<br />

Plots 1 <strong>and</strong> 4 revealed association of infected trees at all distances between all compared dates<br />

with the exception of the 2005-2009 period in Plot 1, where statistically significant association<br />

was observed below 10m only (Figure 4). Plot 2 showed significant association for distances<br />

shorter than 35m for all comparisons. Plot 3 showed significant association for short distances<br />

(20m) for 2003-2004 <strong>and</strong> 2004-2005 <strong>and</strong> for almost all distances for the 2005-2009 comparison.<br />

4. Discussion<br />

As suggested in previous research on this host-pathogen relationship in the same region <strong>and</strong> at<br />

the same scale (Gouveia et al. 2005), the infection pattern of chestnut blight at the orchard level<br />

is r<strong>and</strong>om at the beginning of the disease spread process. It becomes later aggregated when<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.C. Azevedo et al. 2010. Spatial dynamics of chestnut blight disease at the plot level using the Ripley’s K function<br />

459<br />

contamination occurs from the initially infected trees either naturally <strong>and</strong> by means of<br />

management practices, such as pruning, that increases infection at near distances. In this study<br />

we analysed data from a period of time larger than in Gouveia (2005), in a stage of dispersal<br />

when blight is present in the entire region <strong>and</strong> when spread within orchards at short distances<br />

already took place. Therefore, clustered patterns were to be expected for infected trees in all<br />

plots. This happened only in Plots 1 <strong>and</strong> 4, however. Plots 2 <strong>and</strong> 3, showed a pattern generally<br />

r<strong>and</strong>om in 2003, 2004 <strong>and</strong> 2005. The reasons for this are still unknown.<br />

2003 2004<br />

20<br />

20<br />

15<br />

15<br />

10<br />

10<br />

L(t) - t<br />

5<br />

0<br />

L(t) - t<br />

5<br />

0<br />

-5<br />

-5<br />

-10<br />

-10<br />

-15<br />

0 10 20 30 40 50<br />

-15<br />

0 10 20 30 40 50<br />

Distance (m)<br />

Distance (m)<br />

2005 2009<br />

20<br />

20<br />

15<br />

15<br />

10<br />

10<br />

L(t) - t<br />

5<br />

0<br />

L(t) - t<br />

5<br />

0<br />

-5<br />

-5<br />

-10<br />

-10<br />

-15<br />

-15<br />

0 10 20 30 40 50<br />

0 10 20 30 40 50<br />

Distance (m)<br />

Distance (m)<br />

Figure 2. L(t)-t plots (solid lines) for 2003, 2004, 2005 <strong>and</strong> 2009 dead <strong>and</strong> infected chestnut trees in Plot 1.<br />

Dashed lines are 0.025 <strong>and</strong> 0.975 quantiles of L(t) - t estimated from 200 Monte Carlo simulations.<br />

2003 2004<br />

20<br />

20<br />

15<br />

15<br />

10<br />

10<br />

L(t) - t<br />

5<br />

0<br />

L(t) - t<br />

5<br />

0<br />

-5<br />

-5<br />

-10<br />

-10<br />

-15<br />

-15<br />

0 10 20 30 40 50<br />

0 10 20 30 40 50<br />

Distance (m)<br />

Distance (m)<br />

2005 2009<br />

20<br />

20<br />

15<br />

15<br />

10<br />

10<br />

L(t) - t<br />

5<br />

0<br />

L(t) - t<br />

5<br />

0<br />

-5<br />

-5<br />

-10<br />

-10<br />

-15<br />

0 10 20 30 40 50<br />

-15<br />

0 10 20 30 40 50<br />

Distance (m)<br />

Distance (m)<br />

Figure 3. L(t)-t plots (solid lines) for 2003, 2004, 2005 <strong>and</strong> 2009 dead <strong>and</strong> infected chestnut trees in Plot 3.<br />

Dashed lines are 0.025 <strong>and</strong> 0.975 quantiles of L(t) - t estimated from 200 Monte Carlo simulations.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.C. Azevedo et al. 2010. Spatial dynamics of chestnut blight disease at the plot level using the Ripley’s K function<br />

460<br />

It should also be expected that the location of infected trees in one date was associated with the<br />

location of infected tree in the previous date. We observed significant association in most of the<br />

cases, stronger for shorter distances. This seems to corroborate the previously presented<br />

hypothesis, according to which infected trees are spreading blight to the nearer neighbouring<br />

trees. In any case the spread of chestnut blight was very fast at the orchard level. Notice the<br />

infection <strong>and</strong>/or death in the 2005-2009 period (Table 1; Fig 1).<br />

The role of management practices in the spread of the disease is still unclear but it is certain that<br />

fast spread of blight as observed here could be also due to anthropogenic factors such as the<br />

infection of adjacent trees with infected tools.<br />

2003-2004 2004-2005<br />

20<br />

20<br />

15<br />

15<br />

10<br />

10<br />

L(t) - t<br />

5<br />

0<br />

L(t) - t<br />

5<br />

0<br />

-5<br />

-5<br />

-10<br />

-10<br />

-15<br />

0 10 20 30 40 50<br />

-15<br />

0 10 20 30 40 50<br />

Distance (m)<br />

Distance (m)<br />

2005-2009<br />

20<br />

15<br />

10<br />

L(t) - t<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

0 10 20 30 40 50<br />

Distance (m)<br />

Figure 4. L(t)-t plots (solid lines) for comparisons of dead <strong>and</strong> infected trees in Plot 1 for the 2003-2004,<br />

2004-2005 <strong>and</strong> 2005-2009 periods. Dashed lines are 0.025 <strong>and</strong> 0.975 quantiles of L(t) - t estimated from<br />

200 Monte Carlo simulations.<br />

5. Conclusion<br />

In this study we found that in 2 of the 4 studied orchards there was significant clustering of<br />

infected trees in any of the dates observed. In the other two cases infected <strong>and</strong> dead trees<br />

showed a r<strong>and</strong>om pattern. Infected trees in one date were spatially associated with trees infected<br />

the previous date. The results indicate that fast short distance spread of chestnut blight occurs<br />

within orchards.<br />

References<br />

Dixon, P.M., 2002. Ripley’s K function. P. 1796–1803 in Encyclopedia of Environmetrics,<br />

Volume 3, A.H. El-Shaarawi e W.W. Piegorsch (Eds.). John Wiley & Sons, Ltd,<br />

Chichester.<br />

Haase, P., 1995. Spatial pattern analysis in ecology based on Ripley's K-function: introduction<br />

<strong>and</strong> methods of edge correction. Journal of Vegetation Science 6: 575-582.<br />

Getis, A. & Franklin, J., 1987. Second-order neighborhood analysis of mapped point patterns.<br />

Ecology 68: 473-477.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.C. Azevedo et al. 2010. Spatial dynamics of chestnut blight disease at the plot level using the Ripley’s K function<br />

461<br />

Gouveia, E., V. Coelho & J. Azevedo, 2005. Epidemiologia do cancro do castanheiro. Dinâmica<br />

da distribuição espacial de Cryphonectria parasitica (Murrill) Barr. In Páscoa, F. & Silva,<br />

R. (eds.) Actas do 5º Congresso Florestal Nacional, Sociedade Portuguesa de Ciências<br />

Florestais, Lisboa. 12 pp.<br />

Ripley, B.D., 1976. The second order analysis of stationarity processes. Journal of Applied<br />

Probability 13: 255-266.<br />

Ostfeld RS, Glass GE, Keesing F (2005). Spatial epidemiology: an emerging (or re-emerging)<br />

discipline. Trends in Ecology & Evolution 20: 328-336.<br />

Holdenrieder O, Pautasso M, Weisberg PJ, Lonsdale D (2004). Tree diseases <strong>and</strong> l<strong>and</strong>scape<br />

processes: the challenge of l<strong>and</strong>scape pathology. Trends in Ecology & Evolution 19: 446-<br />

452.<br />

Kelly M, Meentemeyer RK (2002). L<strong>and</strong>scape dynamics of the spread of sudden oak death.<br />

Photogramm. Eng. Remote Sens. 68: 1001-1009.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Batista et al. 2010. The third dimension in l<strong>and</strong>scape metrics analysis applied to central Alentejo – Portugal<br />

462<br />

The third dimension in l<strong>and</strong>scape metrics analysis applied to central<br />

Alentejo - Portugal<br />

Teresa Batista 1,3* , Paula Mendes 2 & Luisa Carvalho 3<br />

1<br />

ICAAM, University of Évora, Portugal<br />

2<br />

University of Évora, Portugal<br />

3<br />

Intermunicipal Community of Central Alentejo (CIMAC), Portugal<br />

Abstract<br />

L<strong>and</strong>scape metrics have been widely developed over the last two decades, although the question<br />

remains: How does l<strong>and</strong>scape metrics relates with ecological processes<br />

One of the major recent developments in l<strong>and</strong>scape metrics analysis was the third dimension<br />

integration. Topography has an extremely important role on ecosystems function <strong>and</strong> structure,<br />

even though the common analysis in l<strong>and</strong>scape ecology only conceives planimetric surface<br />

which leads to some erroneous results, particularly in mountain areas.<br />

The analytical process tested patch, class <strong>and</strong> l<strong>and</strong>scape metrics behavior in 11 sample areas of<br />

100 sqkm each in several topographical conditions of Central Alentejo. It is presented the<br />

significance analysis of the results achieved in planimetric <strong>and</strong> 3D environments.<br />

Keywords: L<strong>and</strong>scape metrics; 3D, topography, Local L<strong>and</strong>scape Units, Alentejo, OTALEX<br />

1. Introduction<br />

L<strong>and</strong>scape ecology studies l<strong>and</strong>scape structure, functions <strong>and</strong> changes. L<strong>and</strong>scape structure is<br />

characterized by the composition <strong>and</strong> configuration of l<strong>and</strong>scape patterns. One of the main<br />

premise is that l<strong>and</strong>scape structure is connected with l<strong>and</strong>scape functions <strong>and</strong> processes (Turner<br />

1989, von Drop & Opdam 1987, McIntyre & Wiens 1999). 3D-issue in L<strong>and</strong>scape Ecology<br />

have been studied <strong>and</strong> applied by several researchers in the past 10 years, in many different<br />

approaches (Dorner et al 2002, Bowden et al 2003, Jenness 2004, Lefsky et al 2002, MacNab<br />

1992, Pike 2000, Sebastiá 2004, McGarigal et al 2008). Topography is actually a key factor for<br />

many ecological processes, such as erosion, flow direction <strong>and</strong> accumulation, temperature <strong>and</strong><br />

biodiversity distribution <strong>and</strong> fire (Swanson et al 1988, Burnett et al 1998, Bolstad et al 1998,<br />

Davis & Goetz, 1990 <strong>and</strong> Blaschke et al 2004 ). However it is not taken in to account in most<br />

l<strong>and</strong>scape ecological studies. Only a few recent studies applied 3D to l<strong>and</strong>scape metrics<br />

(Hoechstetter et al 2006, Hoechstetter et al 2008, Hoechstetter 2009, Jenness 2004, Jenness<br />

2010, Walz et al 2010). Others issues like viewsheds <strong>and</strong> l<strong>and</strong>scape preferences have been<br />

studied by Sang et al (2008).<br />

This paper presents part of the l<strong>and</strong>scape studies carried out by the Environmental Indicators<br />

Working Group (EIWG) of OTALEX - Alentejo Extremadura Territorial Observatory<br />

(www.ideotalex.eu) (in OTALEX II Project co-financed by Operational Program for<br />

Cooperation between cross border Regions of Spain <strong>and</strong> Portugal - POCTEP), in Central<br />

Alentejo (Portugal). The main questions are analyzed:<br />

• Are there significant differences in l<strong>and</strong>scape metrics calculated using real surface area<br />

(3D) instead of planimetric area (2D)<br />

• Are there significant differences between the sample areas<br />

* Corresponding author. Tel.:+351266749420 - Fax:+351266749425<br />

Email address: tbatista@cimac.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Batista et al. 2010. The third dimension in l<strong>and</strong>scape metrics analysis applied to central Alentejo – Portugal<br />

463<br />

2. Methodology<br />

2.1 Characterization of study area<br />

The study area is located in Central Alentejo, South of Portugal. It covers about 7.400 sqkm <strong>and</strong><br />

has about 175000 inhabitants, concentrated in small <strong>and</strong> median villages <strong>and</strong> cities. Altimetry<br />

varies between 7 <strong>and</strong> 648 m. We selected 11 sample areas, of 100 sqkm each, located along<br />

Central Alentejo, representing 15% of the total area (figure 1).<br />

2.2 Local L<strong>and</strong>scape Units (LLU)<br />

The definition of l<strong>and</strong>scape units (paths) was based on Corinne L<strong>and</strong> Cover level 5 (CLC N5)<br />

map at scale 1:10.000 (Batista in press), altimetry (MDT 25m) <strong>and</strong> soil units (at sale 1:25.000).<br />

L<strong>and</strong> cover (LC) map applies hierarchical CLC N5 legend developed by Guiomar et al (2006,<br />

2009), with 295 LC classes. The l<strong>and</strong> cover map was elaborated using digital ortophotomaps<br />

from 2005 (from DGRF 2006) <strong>and</strong> field validation at the end of 2008. The LC map has been<br />

previously generalized to create the LLU map. From the overlay of these maps derived 103<br />

Local L<strong>and</strong>scape Units (LLU) (figura 2).<br />

2.3 True Surface Area <strong>and</strong> Perimeter Calculation <strong>and</strong> L<strong>and</strong>scape Metrics<br />

3D applied to l<strong>and</strong>scape metrics implies to calculate those using true surface area <strong>and</strong> perimeter<br />

measurements (Hoechstetter 2009). Surface area provides a better estimate of the l<strong>and</strong> area<br />

available than planimetric area, <strong>and</strong> the ratio of this surface area to planimetric area provides a<br />

useful measure of topographic roughness of the l<strong>and</strong>scape (Jenness 2004).<br />

It was used L<strong>and</strong>Metrics-3D developed by Walz et al (2010), which is an ARCGIS extension<br />

that integrates the available tools for calculating true surface area developed by Jenness (2004,<br />

2010) (http://www.jennessent.com/arcgis/surface_area.htm, last modified April 8, 2010) <strong>and</strong> the<br />

fragstats l<strong>and</strong>scape metrics of McGarigal et al (2002). The application uses a moving window<br />

algorithm <strong>and</strong> estimates the true surface area for each grid cell using a triangulation method<br />

(Figure 3). Each of the triangles is located in three-dimensional space <strong>and</strong> connects the focal cell<br />

with the centre points of adjacent cells. The lengths of the triangle sides <strong>and</strong> the area of each<br />

triangle can easily be calculated by means of the Pythagorean Theorem. The eight resulting<br />

triangles are summed up to produce the total surface area of the underlying cell (for details see<br />

Hoechstetter et al 2008, Hoechstetter 2009 or Jenness 2010).<br />

The analytical process integrates the calculation of Patch, Class <strong>and</strong> L<strong>and</strong>scape metrics for the<br />

11 sample areas, for 2D <strong>and</strong> 3D. The metrics analyzed where: Patch Geometry - Patch Area<br />

(Area) <strong>and</strong> Perimeter (Perim), Shape Metrics - Fractal Dimension (FractDim), Perimeter /Area<br />

Racio (Racio), Shape Index (Shape), Density/Edge Metrics - Edge Density (EdgeDens), Edge<br />

Contrast (Edgecont), number of patches (Numofp), Surface Metrology – Average Roughness<br />

(Avrough) <strong>and</strong> RMS Roughness (RMSrough).<br />

3. Results<br />

The statistical analysis involved 221.382 records generated by 3D-L<strong>and</strong>Metrics software, for the<br />

2 dimensions (2D <strong>and</strong> 3D), 11 sample areas <strong>and</strong> 9 l<strong>and</strong>scape metrics. An ANOVA with multiple<br />

comparing of means (LSD de Fisher method) was run (table 3), resulting in pairwise<br />

comparison, for p


T. Batista et al. 2010. The third dimension in l<strong>and</strong>scape metrics analysis applied to central Alentejo – Portugal<br />

464<br />

analysis. However these results should be carefully interpreted as in previous research<br />

developed by Hoechstetter (2009), certain metrics groups do not reveal a significant difference<br />

between their 2D- <strong>and</strong> 3D-versions (e.g. shape metrics), <strong>and</strong> some of the algorithms (especially<br />

for the distance metrics) involve a considerable computational effort.<br />

Figure 1: Sample areas localization. Central Alentejo – Portugal<br />

Figure 2: Sample areas local l<strong>and</strong>scape units<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Batista et al. 2010. The third dimension in l<strong>and</strong>scape metrics analysis applied to central Alentejo – Portugal<br />

465<br />

Figure 3: Method to determine true surface area <strong>and</strong> true surface perimeter of patches. (figure redrawn<br />

according to Jenness 2004 by Hoechstetter et al 2008).<br />

Table 2: Subject factors for significance analysis<br />

Between-Subjects Factors<br />

Value Label Nº<br />

Dimension<br />

1 2D 110691<br />

2 3D 110691<br />

Area<br />

1 Quadr2 (A1) 18990<br />

2 Quadr4 (A2) 31842<br />

3 Quadr5 (A3) 25578<br />

4 Quadr6 (A4) 20538<br />

5 Quadr7 (A5) 18054<br />

6 Quadr8 (A6) 17892<br />

7 Quadr9 (A7) 22626<br />

8 Quadr10 (A8) 22158<br />

9 Quadr11 (A9) 9882<br />

10 Quadr12 (A10) 19980<br />

11 Quadr13 (A11) 13842<br />

Metrics<br />

1 EgdeDens_LSC 24598<br />

2 AvgRough_LSC 24598<br />

3 RMSRough 24598<br />

4 EdgeCont_LSC 24598<br />

5 Shape_LSC 24598<br />

6 NumOfP_LSC 24598<br />

7 Perim_P 24598<br />

8 Ratio_LSC 24598<br />

9 FractDim_LSC 24598<br />

Table 3: ANOVA results<br />

Df Sum Sq Mean Sq F value Pr(>F)<br />

Dimension 1 4,18E+13 4,18E+13 62459,7 < 2.2e-16 ***<br />

SampleArea 10 1,36E+12 1,36E+11 203,75 < 2.2e-16 ***<br />

Metrics 8 6,23E+14 7,79E+13 116345,7 < 2.2e-16 ***<br />

Dimension/Area 10 7,48E+11 7,48E+10 111,66 < 2.2e-16 ***<br />

Dimension/Metrics 8 8,11E+13 1,01E+13 15147,26 < 2.2e-16 ***<br />

Residuals 221344 1,48E+14 6,70E+08<br />

Signif. Codes: 0'***'; 0,001 '**'; 0,01 '*'; 0,05 '.'; 0,1 ' ' ; 1<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Batista et al. 2010. The third dimension in l<strong>and</strong>scape metrics analysis applied to central Alentejo – Portugal<br />

466<br />

Dependent Variable:Rank of Results<br />

Table 4: Pairwise comparison between 2D <strong>and</strong> 3D<br />

95% Confidence Interval for Difference a<br />

(I) Dimension (J) Dimension Mean Difference (I-J) Std. Error Sig. a Lower Bound Upper Bound<br />

2D 3D -27062,292 * 112,650 ,000 -27283,082 -26841,502<br />

3D 2D 27062,292 * 112,650 ,000 26841,502 27283,082<br />

Based on estimated marginal means<br />

*. The mean difference is significant at the ,05 level.<br />

a. Adjustment for multiple comparisons: Least Significant Difference (equivalent to no adjustments).<br />

Table 5: Results of pairwise comparison between the sample areas<br />

Differes from:<br />

Quadrado 2 – A1 Quadrado 4,6,7,10,11,12,13<br />

Quadrado 4 – A2 Quadrado 2,5,6,7,8,10,11,12,13<br />

Quadrado 5 – A3 Quadrado 4,6,7,9,10,11,12,13<br />

Quadrado 6 – A4 Quadrado 2,4,5,7,8,9,10,12,13<br />

Quadrado 7 – A5 Todas as áreas<br />

Quadrado 8 – A6 Quadrado 4,6,7,10,11,12,13<br />

Quadrado 9 – A7 Quadrado 5,6,7,10,11,12,13<br />

Quadrado 10 – A8 Todas as áreas<br />

Quadrado 11 – A9 Todas as áreas excepto a do quadrado 6<br />

Quadrado 12 – A10 Todas as áreas<br />

Quadrado 13 – A11 Todas as áreas<br />

Acknowledgments<br />

The authors would like to thank to Sebastian Hoechstetter the availability of L<strong>and</strong>Metrics-3D<br />

<strong>and</strong> precious comments, <strong>and</strong> to Nuno Neves the text revision.<br />

References<br />

Batista, T. in press. Carta de uso do solo do Alentejo Central e Concelho de Sousel – Legenda Corine<br />

L<strong>and</strong> Cover Nível 5.<br />

Blaschke, T.; D. Tiede & M. Heurich 2004. 3D-l<strong>and</strong>scape metrics to modelling forest structure <strong>and</strong><br />

diversity based on laser-scanning data. In: Thies, M.; B. Koch; H. Spiecker & H. Weinacker (eds.):<br />

Proceedings of the ISPRS working group VIII/2, „Laser-Scanners for <strong>Forest</strong> <strong>and</strong> L<strong>and</strong>scape<br />

Assessment“, International Society of Photogrammetry <strong>and</strong> Remote Sensing 129-132.<br />

Bolstad, P.V.; W. Swank & J. Vose 1998. Predicting Southern Appalachian overstory vegetation with<br />

digital terrain data. L<strong>and</strong>scape Ecology 13, 271-283.<br />

Bowden, D. C.,G. C.White,A. B. Franklin, <strong>and</strong> J. L. Ganey. 2003. Estimating population size with<br />

correlated sampling unit estimates. Journal of Wildlife Management 67:1–10.<br />

Burnett, M.R.; P.V. August; J.H. Brown Jr. & K.T. Killingbeck 1998. The influence of geomorphological<br />

heterogeneity on biodiversity: I. A patch scale perspective. Conservation Biology 12, 363-370.<br />

Davis, F.W. & S. Goetz 1990. Modeling vegetation pattern using digital terrain data. L<strong>and</strong>scape Ecology<br />

4, 69-80.<br />

Dorner, B.; K. Lertzman & J. Fall 2002. L<strong>and</strong>scape pattern in topographically complex l<strong>and</strong>scapes: issues<br />

<strong>and</strong> techniques for analysis. L<strong>and</strong>scape Ecology 17, 729-743.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


T. Batista et al. 2010. The third dimension in l<strong>and</strong>scape metrics analysis applied to central Alentejo – Portugal<br />

467<br />

Guiomar N., Fern<strong>and</strong>es J.P., Cruz C.S., Batista T. <strong>and</strong> Mateus J. 2006. “Sistemas de classificação e<br />

caracterização do uso e ocupação do solo para zonamento microescalar. Pressupostos para a<br />

adaptação da Legenda CORINE L<strong>and</strong> Cover (Nível 5) à escala 1:10000 e análise comparativa de<br />

sistemas de classificação de uso e ocupação do solo”. Proceedings ESIG2006 15-17. Taguspark.<br />

Oeiras.Portugal<br />

Guiomar N., Batista T., Fern<strong>and</strong>es J.P. <strong>and</strong> Cruz C.S. 2009. Corine L<strong>and</strong> Cover Nível 5 – Contributo para<br />

a carta de Uso do Solo em Portugal Continental. AMDE Edts. Évora. Portugal. 226 pps.<br />

Hargis C.D., Bissonette J.A. <strong>and</strong> David J.L. 1998. The behavior of l<strong>and</strong>scape metrics commonly used in<br />

the study of habitat Fragmentation. L<strong>and</strong>scape Ecology 13: 167–186, 1998.<br />

Hoechstetter S., Xuan Thinh N. <strong>and</strong> Walz U. 2006. 3D-Indices for the Analysis of Spatial Patterns of<br />

L<strong>and</strong>scape Structure Proceedings, InterCarto – InterGIS 12. Berlin 108-118.<br />

Hoechstetter, S., Walz, U., Dang, L.H., Thinh, N.X. 2008 Effects of topography <strong>and</strong> surface roughness in<br />

analyses of l<strong>and</strong>scape structure – A proposal to modify the existing set of l<strong>and</strong>scape metrics<br />

L<strong>and</strong>scape Online 3, 1-14.<br />

Hoechstetter, S. 2009. Enhanced methods for analysing l<strong>and</strong>scape Structure. L<strong>and</strong>scape metrics for<br />

characterising three-dimensional patterns <strong>and</strong> ecological gradients. B<strong>and</strong> 6 der Reihe,<br />

Fernerkundung und angew<strong>and</strong>te Geoinformatik“.Rhombos-Verlag, Berlin. ISBN 978-3-941216-<br />

13-6.<br />

Jenness, J. S. 2004. Calculating l<strong>and</strong>scape surface area from digital elevation models. Wildlife Society<br />

Bulletin. 32(3):829-839<br />

Jenness, J. 2010. Surface Area <strong>and</strong> Ratio for ArcGIS (surface_area.exe) v. 1.0.137. Jenness Enterprises.<br />

Available at: http://www.jennessent.com/arcgis/surface_area.htm.<br />

Lefsky, M.A.; W.B. Cohen; G.P. Parker & D.J. Harding 2002. Lidar remote sensing for ecosystem<br />

studies. BioScience 52, 19-30.<br />

Hobson, R. D. 1972. Chapter 8 - surface roughness in topography: quantitative approach. Pages 221–245<br />

in R. J. Chorley, editor. Spatial analysis in geomorphology. Harper & Row, New York, USA.<br />

Pike, R.J. 2000. Geomorphometry - diversity in quantitative surface analysis. Progress in Physical<br />

Geography 24, 1-20.<br />

McGarigal K., Tagil S. <strong>and</strong> Cushman S.A. 2008. Surface metrics: An alternative to patch metrics for the<br />

quantification of l<strong>and</strong>scape structure.<br />

McGarigal, K.; S.A. Cushman; M.C. Neel & E. Ene 2002. FRAGSTATS: Spatial Pattern Analysis<br />

Program for Categorical Maps - Manual for the computer software program produced by the<br />

authors at the University of Massachusetts, Amherst.<br />

McNab, W.H. 1992. A topographic index to quantify the effect of mesoscale l<strong>and</strong>form on site<br />

productivity. Canadian Journal of <strong>Forest</strong> Research 23, 1100-1107.<br />

Sang N., Miller D. <strong>and</strong> Ode Ð.2008. L<strong>and</strong>scape metrics <strong>and</strong> visual topology in the analysis of l<strong>and</strong>scape<br />

preference. Environment <strong>and</strong> Planning B: Planning <strong>and</strong> Design, volume 35, pages 504-520.<br />

Sebastiá, M.-T. 2004. Role of topography <strong>and</strong> soils in grassl<strong>and</strong> structuring at the l<strong>and</strong>scape <strong>and</strong><br />

community scales. Basic <strong>and</strong> Applied Ecology 5, 331-346.<br />

Swanson, F.J.; T.K. Kratz; N. Caine & R.G. Woodmansee 1988. L<strong>and</strong>form effects on ecosystem patterns<br />

<strong>and</strong> processes. BioScience 38, 92-98.<br />

Turner, M.G. 1989. L<strong>and</strong>scape ecology: the effect of pattern on process. Annual review of ecology <strong>and</strong><br />

systematics 20, 171-197.<br />

van Dorp, D. & P.F.M. Opdam 1987. Effects of patch size, isolation <strong>and</strong> regional abundance on forest<br />

bird communities. L<strong>and</strong>scape Ecology 1, 59-73.<br />

Walz U., Hoechstetter S., Vock D., Dang L.H.2010. L<strong>and</strong>Metrics-3D. L<strong>and</strong>scape metrics for raster-based<br />

structureanalysis in three dimensions. Leibniz Institute of Ecological <strong>and</strong> Regional Development<br />

(IOER). Weberplatz 1, D-01217 Dresden, Germany. 10 pps.<br />

McIntyre N.E. & Wiens J.A. 1999. Interactions between l<strong>and</strong>scape structure <strong>and</strong> animal behavior: the roles of<br />

heterogeneously distributed resources <strong>and</strong> food deprivation on movement patterns. L<strong>and</strong>scape Ecology 14:<br />

437-447.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


V. Caboun 2010. New classification <strong>and</strong> utilization of forest functions in l<strong>and</strong>scape<br />

468<br />

New classification <strong>and</strong> utilization of forest functions in l<strong>and</strong>scape<br />

Vladimir Caboun *<br />

National <strong>Forest</strong> Centre – <strong>Forest</strong> Research Institute in Zvolen, T.G. Masaryka 22,<br />

960 92 Zvolen, Slovak Republic<br />

Abstract<br />

Main aim of the research task is scientific assessment of acquired knowledge on functional<br />

effects of forests under real ecological, forest management <strong>and</strong> socio-economic conditions of<br />

the regions of Slovakia. Constructed is new classification system in which we divided strictly<br />

tree species functions underst<strong>and</strong>ing how influences or effects on particular compounds of<br />

environment (- ecological sphere) <strong>and</strong> their utilization by a human society in economic <strong>and</strong><br />

social sphere. In this way can be offered integrated functions of tree species <strong>and</strong> their<br />

communities how goods or service.<br />

To the forest function (tree species function) <strong>and</strong> its classification we access individual. Here<br />

are important indicators structure <strong>and</strong> ecological stability of tree species community. To the<br />

possibilities of utilization of forest functions we use integration of forest functions <strong>and</strong><br />

economic access.<br />

Current ecological or ecosystem approach to the forest functions <strong>and</strong> possibilities of their<br />

utilization changed actual access to forest functions.<br />

Key words: <strong>Forest</strong> functions, classification of forest functions, utilization of the functions of<br />

forest tree species<br />

1. Introduction<br />

<strong>Forest</strong>s <strong>and</strong> other communities of tree species play irreplaceable functions in the l<strong>and</strong>scape from<br />

the viewpoint of the ecological stability of the l<strong>and</strong>scape, its rational utilization <strong>and</strong> sustainable<br />

development. <strong>Forest</strong>s represent a basic l<strong>and</strong>scape forming <strong>and</strong> ecological <strong>and</strong> stabilizing<br />

element of the l<strong>and</strong>scape. They are the most important source of renewable resources <strong>and</strong> thanks<br />

to their functions they play an important role also in the formation <strong>and</strong> protection of individual<br />

components of natural environment as well as the environment changed by anthropogenic<br />

activities <strong>and</strong> anthropic (artificial environment created by a man).<br />

1. 1 Aim of scientific-research activities<br />

Main aim of the research task is to assess recent knowledge on functional effects of forests in<br />

real ecological, forest management <strong>and</strong> socio-economic conditions of individual regions of<br />

Slovakia with the utilization of the latest knowledge of present ecology <strong>and</strong> economics of<br />

natural resources. On the basis of that was created new classification, a classification system,<br />

methodology of the valuation of forest functions <strong>and</strong> it will be proposed the methodology of<br />

determining the rate of ecological-stabilization effect of forests in the l<strong>and</strong>scape.<br />

Main objectives of research task dealing with forest functionsare dissemination of scientific<br />

knowledge on forest functions <strong>and</strong> possibilities of their utilization in the l<strong>and</strong>scape, construction<br />

of classification system of forest functions, construction of a system of assessment <strong>and</strong><br />

valuation of forest functions <strong>and</strong> tree species communities from the viewpoint of their<br />

multifunctional utilization.<br />

* Email address: caboun@nlcsk.org<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> - New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape<br />

Ecology International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Portugal.


V. Caboun 2010. New classification <strong>and</strong> utilization of forest functions in l<strong>and</strong>scape<br />

469<br />

1.2 Significance of the issues solved<br />

The importance of these issues follows also from the fact the European Commission issued<br />

COM no. (88) 255 concerning the strategy <strong>and</strong> action plan of the Community in forestry <strong>and</strong> set<br />

up in total 6 objectives for forest sector, of them 4 are directly related to the solved issue:<br />

• Support the participation of the whole forest sector in planning the utilization of the<br />

l<strong>and</strong>scape <strong>and</strong> thus to contribute to rural development<br />

• Contribute to the protection <strong>and</strong> improvement of the environment<br />

• Secure dynamic development of forestry that would enable better fulfilment of<br />

individual forest functions<br />

• Enhance importance of forests as a natural environment for recreation.<br />

It follows from the mentioned above that the future of forestry depends on the importance of the<br />

forests in a society. Interrelations of human society <strong>and</strong> tree species <strong>and</strong> their utilization of their<br />

functions have changed in time <strong>and</strong> space. A man used the functions of tree species <strong>and</strong> their<br />

communities in the l<strong>and</strong>scape in dependence on the number of a concrete human population,<br />

natural conditions, <strong>and</strong> way of living as well as in dependence on social, economic <strong>and</strong> cultural<br />

development of a society.<br />

In accordance with EU forestry strategy one of basic goals of forest policy in Slovakia is<br />

enhancement of multifunctional (functionally integrated) management of forests <strong>and</strong> protection<br />

of the potential of their functions. We must h<strong>and</strong>le functional potential of forests as the natural<br />

wealth <strong>and</strong> to preserve <strong>and</strong> improve it by proper management.<br />

Among the most serious problems limiting effective applying the system of multifunctional<br />

forest management is mainly discordance between social order for forest functions <strong>and</strong> their<br />

economic funding.<br />

2. Methodology<br />

2.1 Theoretical <strong>and</strong> methodical starting points<br />

Despite the fact the issues of forest functions were solved mainly in the 70-80s of the past<br />

century the solution has not been completely <strong>and</strong> satisfactorily finished what concerns the<br />

functions of tree species <strong>and</strong> their communities in new ecological <strong>and</strong> socio-economic<br />

conditions of Slovakia. Recently prevailing perception of the nature <strong>and</strong> forest, which served the<br />

man <strong>and</strong> his requirements caused that forest functions were considered services with purposeful<br />

selection <strong>and</strong> social utilitarian prioritisation.<br />

Modern ecological approach to forest <strong>and</strong> forest functions in the l<strong>and</strong>scape must consider the<br />

latest knowledge on ecosystem research of forest. This view at forest ecosystems must<br />

necessarily consider long-term time factor bringing about various dynamics of the changes of<br />

ecological, economic <strong>and</strong> social conditions, but mainly different view at forest functions <strong>and</strong><br />

their utilization. From this viewpoint the way of functional integration seems to be substantially<br />

more effective <strong>and</strong> more pragmatic than the way of purposeful differentiation <strong>and</strong> prioritisation<br />

of some of the functions.<br />

To be able to use this approach it is necessary to extend greatly scientific knowledge on forest<br />

functions <strong>and</strong> possibilities of their utilization in the l<strong>and</strong>scape as well as to construction a new<br />

classification system of forest functions that would consider ecological approach to forest as to<br />

ecosystem.<br />

This approach presupposes construction of basic typology <strong>and</strong> the system of the evaluation of<br />

forest functions potential <strong>and</strong> assessment of real fulfilment of the functions by forest growing<br />

under various site conditions, various types of the l<strong>and</strong>scape (with various use <strong>and</strong> degree of<br />

anthropic changes), with regard to the health condition of a real forest, its current tree species<br />

composition, age <strong>and</strong> spatial arrangement of forest as well as to with regard to its ecological<br />

stability considering expected global <strong>and</strong> regional (mainly climatic) changes with regard to<br />

social requirements <strong>and</strong> the interests of forest owners.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> - New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape<br />

Ecology International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Portugal.


V. Caboun 2010. New classification <strong>and</strong> utilization of forest functions in l<strong>and</strong>scape<br />

470<br />

<strong>Forest</strong> management as a production sector lives on the sale of own products. From this<br />

viewpoint production forest functions brings profit <strong>and</strong> all other forest functions are<br />

only a load for forest manager, it means they are not equal to production function. The<br />

core of the integration of forest functions is namely mutual comparison <strong>and</strong> evaluation<br />

of various forest functions, their reflection in the system of management in forest <strong>and</strong><br />

assessment of benefits resulting from various ways <strong>and</strong> interlinking of forest functions<br />

use into optimal proportions. <strong>Forest</strong> manager must know which forest benefits the<br />

society needs to be able to set the goals of management.<br />

Our task was not simple as it is to construct the classification system of the assessment of the<br />

potential of forest functions <strong>and</strong> real fulfilment of the functions by forest growing in various site<br />

conditions <strong>and</strong> types of the l<strong>and</strong>scape with various utilization <strong>and</strong> degree of anthropic changes<br />

with regard to real state of forest, its current tree species composition, age <strong>and</strong> spatial structure,<br />

ecological stability considering not only historical development <strong>and</strong> present state but also<br />

expected global <strong>and</strong> regional (mainly climatic) changes <strong>and</strong> anthropogenic effects as well as<br />

with regard to social requirements <strong>and</strong> interests of the owners.<br />

2.2 Analysis of ecological-stabilization <strong>and</strong> functional effectiveness of forest<br />

ecosystems in the l<strong>and</strong>scape<br />

On the basis of available literature experimental results there was carried out primary analysis of<br />

the functional effectiveness of forest ecosystems in the l<strong>and</strong>scape <strong>and</strong> the system for its<br />

detection <strong>and</strong> classification was worked out. This system follows up the system of the<br />

classification of ecological stability (Caboun 2002, 2003), as long-term ecological stability is a<br />

basic precondition for securing long-term functionality of forests.<br />

We underst<strong>and</strong> ecological stability as an ability of the ecosystem to resist or compensate<br />

external as well as internal effects without any marked permanent disturbing of the functional<br />

structure of this system.<br />

Natural ecosystem develops in accordance with given conditions <strong>and</strong> usual abiotic <strong>and</strong> biotic<br />

factors. These conditions <strong>and</strong> factors form ecosystem (the effect of the environment) what<br />

appears also for the given conditions in its specific structure (tree species composition, age <strong>and</strong><br />

spatial structure) <strong>and</strong> subsequently in its ecological stability. Optimal solution from the<br />

viewpoint of ecological stability, <strong>and</strong> thus also optimal functionality of the ecosystem is on the<br />

basis of our knowledge solution of nature through natural ecosystems. A man from the<br />

viewpoint of the need of satisfying own needs influenced the structure of forests to different<br />

extent, <strong>and</strong> thus he influenced also their ecological balance, ecological stability <strong>and</strong><br />

subsequently resulting fulfilment of the forest functions.<br />

Graphs of percent reduction of partial ecological stability in dependence on the degree of<br />

difference of the studied indicator of a real (assessed) forest ecosystem in comparison with<br />

optimal forest ecosystem corresponding to the site are a part of the classification system of<br />

partial ecological stability of individual indicators. Construction of models or their specification<br />

to the level of forest types or types of forest is a dem<strong>and</strong>ing <strong>and</strong> long-term task of further<br />

research in cooperation with the people who implemented <strong>and</strong> verified the proposed system.<br />

The determination of the ecological stability for individual time horizons is based on individual<br />

development phases or their changes during the studied period as well as presupposed site<br />

changes during this time.<br />

For each development phase it is possible to determine in general the range of its primary –<br />

initial ecological stability on the basis of hypothetical models of ecological stability <strong>and</strong> its<br />

components of individual development phases of tree species.<br />

The sense <strong>and</strong> practical importance of ecological stability lies in the fact that on the basis of<br />

found facts <strong>and</strong> values it is possible to propose <strong>and</strong> optimal way of management in accordance<br />

with natural regularities in a way to strengthen the required component of ecological stability –<br />

resistance or flexibility with regard to the fulfilment of the required forest functions, the time of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> - New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape<br />

Ecology International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Portugal.


V. Caboun 2010. New classification <strong>and</strong> utilization of forest functions in l<strong>and</strong>scape<br />

471<br />

fulfilment of these functions <strong>and</strong> mainly with regard to expected decisive factors influencing the<br />

existence <strong>and</strong> ecological stability of a concrete forest. Interlinking of functional effectiveness<br />

<strong>and</strong> ecological stability through the structure corresponding to site follows up with the proposal<br />

of starting points for the construction of the classification system of forest functions.<br />

3. Results: <strong>Forest</strong> functions – classification <strong>and</strong> possibilities of their utilization<br />

In the proposal of the classification system of forest functions there are clearly distinct forest<br />

functions being perceived as the effect of forest on individual components of the environment<br />

from the utilization of these functions by a man. Systematic solution of the methodological<br />

approach to forest functions <strong>and</strong> their classification is presented in Figure 1.<br />

Figure 1 Ecosystem approach to forest <strong>and</strong> other communities of tree species in the l<strong>and</strong>scape, their<br />

functions <strong>and</strong> possibilities of functions utilization in economic <strong>and</strong> social fields (Caboun 2005)<br />

We distinguish basic forest functions affecting abiotic components of the environment (air,<br />

water, soil) <strong>and</strong> biotic components (plants, animals, microorganisms, man).<br />

In this way tree species <strong>and</strong> their communities fulfil in the l<strong>and</strong>scape edaphic, atmospheric,<br />

hydric <strong>and</strong> lithic function what concerns abiotic components of the ecosystem, <strong>and</strong> phytobiotic,<br />

zoobiotic, microbiotic <strong>and</strong> anthropic function what concerns biotic components of the<br />

ecosystem. In other words it is the quality <strong>and</strong> quantity of the effect of tree species <strong>and</strong> their<br />

communities on the soil, climate, water, rocks, plants, animals, microorganisms <strong>and</strong> man.<br />

These functions are divided into partial functions. For example edaphic function comprises soil<br />

forming, soil reclamation <strong>and</strong> soil protection functions, which consists of erosion control, anti<br />

deflation, anti slides function, avalanche control <strong>and</strong> bank protection function.<br />

A human society may use a complex of these functions for economic purposes or in a social<br />

area. Then forestry, water management, game management, agriculture, energy industry, food<br />

industry, building industry, chemical industry, cosmetics, pharmacy etc. belong to anthropic<br />

fields using forest functions in economic area. Similarly, forest functions may be used in social<br />

area, it means for recreation, curing, hygiene, for nature protection, formation <strong>and</strong> protection of<br />

the environment, science <strong>and</strong> research, education <strong>and</strong> training, aesthetics <strong>and</strong> arts, culture <strong>and</strong><br />

history <strong>and</strong> others.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> - New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape<br />

Ecology International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Portugal.


V. Caboun 2010. New classification <strong>and</strong> utilization of forest functions in l<strong>and</strong>scape<br />

472<br />

This classification of forest functions creates a basic information base for the possibility of the<br />

utilization of the functions of tree species <strong>and</strong> their communities in the l<strong>and</strong>scape by human<br />

society.<br />

The aim is to construct the classification system for the assessment of the potential of forest<br />

functions <strong>and</strong> real fulfilment of the functions of a forest growing in different site conditions <strong>and</strong><br />

types of the l<strong>and</strong>scape with various use <strong>and</strong> degree of anthropic changes. An emphasis will be<br />

put on the real state of forest, its current tree species composition, age <strong>and</strong> spatial structure,<br />

ecological stability considering not only historical development but also present state as well as<br />

its expected development, global <strong>and</strong> regional (mainly climatic) changes <strong>and</strong> anthropogenic<br />

effects, <strong>and</strong> social requirements <strong>and</strong> the interests of the forests owners.<br />

A new important element in the utilization of forest functions is financial reimbursement paid to<br />

the forest owner for the provided services<br />

Philosophy of practical use <strong>and</strong> methodology in detecting, classifying <strong>and</strong> valuation of forest<br />

functions in the l<strong>and</strong>scape is expressed briefly in following points:<br />

• Setting apart a part of the l<strong>and</strong>scape for the assessment <strong>and</strong> valuation of forest functions<br />

• Ecological-functional typifying of determined part of the l<strong>and</strong>scape <strong>and</strong> attributing of<br />

corresponding (potential) communities of tree species<br />

• Evaluation of the difference of the structure of real forests with optimal structure of<br />

potential forests in the determined part of the l<strong>and</strong>scape<br />

• Determination of ecological-stabilization rate (effect) of real forests in the determined part<br />

of the l<strong>and</strong>scape (classification of ecological stability of forests <strong>and</strong> particular part of the<br />

l<strong>and</strong>scape)<br />

• Assessment of real functionality of forests <strong>and</strong> communities of tree species in the<br />

determined part of the l<strong>and</strong>scape from the viewpoint of their structure<br />

• Assessment of social requirements on the use of the determined part of the l<strong>and</strong>scape <strong>and</strong> on<br />

fulfilment of the functions by forest tree species <strong>and</strong> their communities<br />

• Appraisal <strong>and</strong> valuation of forest functions with regard to the type of the determined part of<br />

the l<strong>and</strong>scape <strong>and</strong> requirements on fulfilment <strong>and</strong> utilization of forest functions in the<br />

determined part of the l<strong>and</strong>scape<br />

• Proposal of management <strong>and</strong> measures to optimise the structure of forests <strong>and</strong> their<br />

functions in the determined part of the l<strong>and</strong>scape with regard to the state, ecological stability<br />

<strong>and</strong> financially grounded requirements on the use of forest functions <strong>and</strong> the functions of<br />

communities of tree species in this particular part of the l<strong>and</strong>scape.<br />

From the viewpoint of prediction of the development of fulfilment <strong>and</strong> utilization of the<br />

functions of tree species communities in the l<strong>and</strong>scape it must be noted that as it is possible with<br />

ecological stability to determine its probable development on the basis of supposed changed of<br />

site conditions <strong>and</strong> the structure of forest ecosystem, it is also possible to predict the<br />

development of the capability of this ecosystem to fulfil individual functions in the l<strong>and</strong>scape or<br />

environment. But it is very difficult to predict the need, capability <strong>and</strong> social willingness of the<br />

utilization of these functions with their adequate financial reimbursement.<br />

It is possible <strong>and</strong> appropriate to present in graph the comparison of potential <strong>and</strong> present<br />

fulfilment of the functions by tree species <strong>and</strong> their communities in the studied area.<br />

Similarly, real utilization of the functions of tree species <strong>and</strong> their communities on the studied<br />

territory as well as comparison of real utilization of the functions with social order can be<br />

illustrated in graph.<br />

4. .Discussion <strong>and</strong> conclusion<br />

From insinuated way is possible to compare real possibilities of particular ecosystem to fulfil<br />

required functions what subsequently shows the need of the management of this ecosystem. The<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> - New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape<br />

Ecology International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Portugal.


V. Caboun 2010. New classification <strong>and</strong> utilization of forest functions in l<strong>and</strong>scape<br />

473<br />

management comprises influencing the structure of community <strong>and</strong> thus also ecological stability<br />

<strong>and</strong> fulfilment of individual functions. With regard to the fact that in our solution we prefer<br />

integration of functions <strong>and</strong> not their prioritisation, a more complex utilization of forest<br />

functions will be aimed at close to nature management, potential – optimal forest community.<br />

The presented approach has not only maximal economic benefit but ecological stability of<br />

concrete ecosystem as a part of the l<strong>and</strong>scape where it is located is increasing <strong>and</strong> the<br />

importance of tree species <strong>and</strong> their communities, mainly of forest in the l<strong>and</strong>scape, will<br />

increase substantially as well.<br />

The core <strong>and</strong> substance of the integration of forest functions is namely mutual comparison <strong>and</strong><br />

evaluation of various forest functions, their reflection in the system of management in forest <strong>and</strong><br />

considering benefits following from various ways <strong>and</strong> degrees of interlinking of forest functions<br />

<strong>and</strong> their utilization into optimal proportions. <strong>Forest</strong> manager must know what forest benefits<br />

the society needs to be able to set properly the objectives of management in the given area that<br />

will secure optimal utilization of forest functions with their concrete structure <strong>and</strong> with regard to<br />

their ecological stability for clearly defined time period.<br />

Then priorities will follow from the proposal of the management <strong>and</strong> measures for optimisation<br />

of the structure of forests <strong>and</strong> their functions in determined part of the l<strong>and</strong>scape with regard to<br />

the state, ecological stability <strong>and</strong> financially reasoned requirements on the utilization of the<br />

functions of forests <strong>and</strong> tree species communities in this l<strong>and</strong>scape.<br />

On the basis of current knowledge <strong>and</strong> the latest approaches to forest functions, functions of<br />

forest tree species <strong>and</strong> their communities the way of functional integration seems to be more<br />

effective <strong>and</strong> more pragmatic than the way of purposeful differentiation <strong>and</strong> prioritisation of<br />

some of the functions.<br />

Extending of scientific knowledge on forest functions, forest tree species <strong>and</strong> their communities<br />

<strong>and</strong> possibilities of their use in the l<strong>and</strong>scape will enable not only their real use in the<br />

environment but also construction of a new classification system of the functions of forests,<br />

forest tree species <strong>and</strong> their communities considering ecological <strong>and</strong> subsequently economic<br />

approach.<br />

This approach presupposes construction of basic typology <strong>and</strong> the system of the assessment of<br />

forest functions potential as well as the assessment of real fulfilment of the functions of forest<br />

growing in various site conditions, various types of the l<strong>and</strong>scape (various use <strong>and</strong> degree of<br />

anthropic changes), with regard to the health condition of real forest, its present tree species<br />

composition, age <strong>and</strong> spatial structure, its ecological stability that considers expected global <strong>and</strong><br />

regional (mainly climatic) changes with regard to social requirements <strong>and</strong> the interests of the<br />

owners of forests.<br />

References<br />

Caboun, V., 2002. System of indicators of forest ecological stability <strong>and</strong> its classification.<br />

Proceedings from the international scientific symposium on “New trends in detecting <strong>and</strong><br />

monitoring the state of forest”, Technical University Zvolen, p. 116-135, in Slovak<br />

Caboun, v., 2003. Classification of ecological equilibrium <strong>and</strong> ecological stability on an<br />

example of a model territory. Ecological research <strong>and</strong> protection of the nature of the<br />

Carpathian Mts. Proceedings of the international scientific conference, TU Zvolen,<br />

Lesoprojekt Zvolen, Slovak Ecological Society at SAV (Slovak Academy of Sciences),<br />

2003, p. 134 – 140.<br />

Caboun, v., 2005. Spatial arrangement of the territory – determining of ecological– functional<br />

areas within the project “Revitalization of forest ecosystems on the territory of the High<br />

Tatra Mts. affected by wind calamity on 19.11.2004. Ecological Studies VI.<br />

Metamorphoses of the nature protection in the Tatra Mts. Slovak Ecological Society at<br />

SAV (Slovak Academy of Sciences), ISBN 80-968901-1-3-1, p. 126-136.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong> - New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape<br />

Ecology International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Portugal.


A.M. Carvalho et al. 2010. Connecting l<strong>and</strong>scape conservation <strong>and</strong> management with traditional ecological knowledge<br />

474<br />

Connecting l<strong>and</strong>scape conservation <strong>and</strong> management with traditional<br />

ecological knowledge: does it matter how people perceive l<strong>and</strong>scape<br />

<strong>and</strong> nature<br />

Ana M. Carvalho 1* , Margarida T. Ramos 2 & Amélia Frazão-Moreira 3<br />

1<br />

CIMO/<strong>ESA</strong>B, Instituto Politécnico de Bragança, Portugal<br />

2<br />

Projecto Cultibos, yerbas i saberes, FRAUGA <strong>and</strong> IPB, Portugal<br />

3<br />

CRIA/Faculdade Ciências Sociais e Humanas da Universidade Nova de Lisboa,<br />

Portugal<br />

Abstract<br />

Ethnobotanical surveys conducted in Trás-os-Montes (Portugal) highlighted a renewed interest<br />

in cultural values of l<strong>and</strong>scapes. Long term interactions between traditional ecological<br />

knowledge (TEK) <strong>and</strong> natural processes provided l<strong>and</strong>scapes characterized by high diversity<br />

<strong>and</strong> relative stability. Rural contexts are facing social <strong>and</strong> economical constraints <strong>and</strong><br />

l<strong>and</strong>scapes are change accordingly.<br />

On a basis of ethnographic methodologies (consented interviews <strong>and</strong> participant observation),<br />

recent l<strong>and</strong>scape changes at a local level <strong>and</strong> people’ perceptions are briefly described <strong>and</strong><br />

discussed as important tools for l<strong>and</strong>scape conservation <strong>and</strong> management.<br />

Young <strong>and</strong> some middle aged people value some of these changes, which they consider less<br />

hard-working <strong>and</strong> a symbol of modernity. Others see actual transformations as a waste of<br />

resources <strong>and</strong> ab<strong>and</strong>onment <strong>and</strong> thus l<strong>and</strong>scape is perceived as unproductive, which is<br />

considered reprehensible. Most of the informants are aware of a dynamic process taking place<br />

<strong>and</strong> conscious that l<strong>and</strong>scape, like themselves, must adapt to changing times.<br />

Keywords: TEK, Portuguese ethnobotany, cultural l<strong>and</strong>scapes, rural l<strong>and</strong>scapes dynamics<br />

1. Introduction: Cultural l<strong>and</strong>scape <strong>and</strong> traditional ecological knowledge (TEK)<br />

Traditional rural l<strong>and</strong>scapes are culturally relevant <strong>and</strong> a consequence of an integration of<br />

natural <strong>and</strong> anthropogenic processes that resulted in a great diversity of sustainable l<strong>and</strong>scapes<br />

(Council of Europe 2000; Antrop 2005).<br />

Human influence, especially related to agriculture <strong>and</strong> l<strong>and</strong>-use patterns, have determined<br />

greater impacts on rural l<strong>and</strong>scape than the ecological features <strong>and</strong> processes over the last<br />

decades. In Europe, long <strong>and</strong> complex history of l<strong>and</strong> uses have promoted a rich diversity of<br />

cultural l<strong>and</strong>scapes that have been shaped by local practices, believes <strong>and</strong> specific purposes, <strong>and</strong><br />

maintained in those areas where physical, socio-economic <strong>and</strong> political constrains have<br />

prevented modernization <strong>and</strong> changing in farming systems until recent times (Vos <strong>and</strong> Meekes<br />

1999; Antrop 2005; Plieninger, Hochtl <strong>and</strong> Spek 2006; Calvo-Iglesias, Fra-Palelo <strong>and</strong> Diaz-<br />

Varela 2009).<br />

Ethnobotanical studies deal with local people knowledge <strong>and</strong> perceptions of nature <strong>and</strong><br />

environment. Traditional ecological knowledge (TEK) has great cultural significance <strong>and</strong> refers<br />

to the use of many wild or domesticated resources <strong>and</strong> the management of natural habitats <strong>and</strong><br />

* Corresponding author. Tel.: +351 273324524<br />

Email address: anacarv@ipb,pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Carvalho et al. 2010. Connecting l<strong>and</strong>scape conservation <strong>and</strong> management with traditional ecological knowledge<br />

475<br />

agroecosystems. TEK refers, as well, to some other important rural activities <strong>and</strong> practices, such<br />

as cattle transhumance, agricultural techniques (e.g. crop rotation, irrigation methods, multi use<br />

parcels <strong>and</strong> partial harvest), l<strong>and</strong> management (e.g. l<strong>and</strong> holding fragmentation, terraces, natural<br />

or artificial boundaries), rituals <strong>and</strong> ceremonies, oral traditions <strong>and</strong> symbolism, communitarian<br />

features <strong>and</strong> settlement patterns (Martin 1995; Cunningham 2001).<br />

Coherent relations between the physical environment, TEK <strong>and</strong> local adaptation result in wellestablished<br />

regionally differentiated patterns of settlements, l<strong>and</strong> use systems <strong>and</strong> field structure<br />

that characterize European cultural l<strong>and</strong>scapes (Vos <strong>and</strong> Meekes 1999). Local people<br />

knowledge <strong>and</strong> traditional rural l<strong>and</strong>scapes are fundamental for the conservation of biodiversity<br />

<strong>and</strong> also a source of information on the past <strong>and</strong> present cultural l<strong>and</strong>scapes, focusing on the<br />

l<strong>and</strong>-use system, management techniques, cultural heritage, <strong>and</strong> farmers’ perception of changes<br />

(Antrop 2005; Calvo-Iglesias, Fra-Palelo <strong>and</strong> Diaz-Varela 2009).<br />

Cultural l<strong>and</strong>scapes dynamics pressuposes change according to changing TEK, values <strong>and</strong><br />

policies. For many centuries these changes had local impact <strong>and</strong> thus cultural l<strong>and</strong>scapes were<br />

perceiveid as rather stable (Vos <strong>and</strong> Meekes 1999; Antrop 2005; Calvo-Iglesias, Crecente-<br />

Maseda, Fra-Paleo, 2006; Carvalho 2010).<br />

Nowadays, rural contexts face sudden <strong>and</strong> faster social <strong>and</strong> economical transformations <strong>and</strong><br />

rural l<strong>and</strong>scapes are rapidly changing <strong>and</strong> diversifying. The geographical <strong>and</strong> social conditions<br />

that led to the isolation of some European regions are no longer prevalent, <strong>and</strong> so changes in the<br />

cultural, economic <strong>and</strong> political contexts of plant use <strong>and</strong> l<strong>and</strong>scape are coming faster <strong>and</strong> faster.<br />

The current reduction in the human population, due to out-migration <strong>and</strong> a general drop in the<br />

birth rate, <strong>and</strong> the ab<strong>and</strong>onment of agriculture have been critical for rural areas <strong>and</strong> ways of life<br />

<strong>and</strong> have promoted the loss of cultural traditions. These changes are affecting the system of<br />

local knowledge of plant resources <strong>and</strong> the maintenance of traditional plant use practices<br />

(Carvalho <strong>and</strong> Morales 2010). L<strong>and</strong>scape changes observed <strong>and</strong> locally perceived are some of<br />

the topics addressed in this paper, based on research in such a region of northeastern Portugal.<br />

2. Methodology<br />

The l<strong>and</strong>scape topic emerged from several ethnobotanical surveys carried out in Trás-os-Montes<br />

(a Portuguese region) for almost nine years (2000-2009) within the scope of three different<br />

research projects that aimed to record <strong>and</strong> document traditional knowledge on plant resources<br />

use, related technologies <strong>and</strong> management.(Ramos, 2008; Frazão-Moreira, Carvalho <strong>and</strong><br />

Martins 2009; Carvalho 2010).<br />

The study area is located in the most northeastern part of the Trás-os-Montes region <strong>and</strong><br />

included in two important natural protected areas (the Natural Park of Montesinho <strong>and</strong> the<br />

Natural Park of Douro Internacional) corresponding to Vinhais, Bragança <strong>and</strong> Mir<strong>and</strong>a do<br />

Douro municipalities. Mostly a mountainous <strong>and</strong> very isolated rural area, with small villages<br />

(many of them less than 100 inhabitants) scattered all over the l<strong>and</strong>scape. The local economy<br />

was/is based on small farming systems, with an important crop production diversity, a high<br />

level of subsistence strategies avoiding productive risks, <strong>and</strong> mostly affected by agriculture<br />

ab<strong>and</strong>onment <strong>and</strong> both population ageing <strong>and</strong> erosion, due to several migratory flows.<br />

Traditional l<strong>and</strong>scapes are characterized by a mosaic composed of different patches finely<br />

linked to each other, particularly highlighted by the seasonal contrasts of the vegetation <strong>and</strong><br />

agricultural activities (e.g. fallows, manure, hay or grazed meadows, orchards, gardens).<br />

Within the research projects, we used a r<strong>and</strong>om stratified sampling of approximately 40% of the<br />

villages in the study area which had a history of agropastoral activities <strong>and</strong> homegardens until<br />

very recently (at least 2005). In every case study, consented semi-structured interviews as well<br />

as participant observation were conducted during all seasons of the year. Informants (a total of<br />

165) were selected using r<strong>and</strong>om sample <strong>and</strong> snow-ball methods (Martin 1995; Alexiades 1996).<br />

In-depth interviews have been held with 30 local experts or key informants (informants with<br />

profound knowledge of a particular aspect of local culture, e.g. shepherds, smugglers, hunters,<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Carvalho et al. 2010. Connecting l<strong>and</strong>scape conservation <strong>and</strong> management with traditional ecological knowledge<br />

476<br />

healers), a sub-group selected from those informants considered knowledgeable by their<br />

neighbors (Martin 1995; Carvalho 2010).<br />

For the purpose of this paper we have only considered the information provided by key<br />

informants (18 women <strong>and</strong> 12 men, nearly all over 60 years old) that have lived most of their<br />

lives in the selected villages, were acquainted with forestry, animal husb<strong>and</strong>ry, agricultural<br />

practices <strong>and</strong> local farming systems <strong>and</strong> culture, <strong>and</strong> that were able to remember or have<br />

participated in different management scenarios of natural <strong>and</strong> traditional l<strong>and</strong>scape (e.g. plows,<br />

clearings, common l<strong>and</strong>s, forestation, road system, irrigation canals <strong>and</strong> mining, for instance).<br />

As a photographic collection <strong>and</strong> an ethnobotanical database were created, it was possible to<br />

compare structural components of traditional l<strong>and</strong>scapes <strong>and</strong> some elements such as l<strong>and</strong> cover,<br />

building <strong>and</strong> infrastructural construction for a time period of almost a decade.<br />

3. Results<br />

A descriptive <strong>and</strong> qualitative analysis of the reported data show that l<strong>and</strong>scape changing in<br />

structure <strong>and</strong> l<strong>and</strong> use has been locally detected in the last two decades in all the studied areas.<br />

Key informants considered that the main signs of l<strong>and</strong>scape structural changing are:<br />

• The emergence of a road system since the 1990s allowing a better physical communication<br />

between some villages <strong>and</strong> nearest towns, as it was very inefficient or nonexistent in many of<br />

cases;<br />

• The village planning since 1975. For instance, sewage system, water distribution network,<br />

paving, small medical centers, rebuilding <strong>and</strong> scattered secondary housing, formal meeting<br />

places;<br />

• Rebuilding <strong>and</strong> preservation of collective facilities (e.g. water mill, forge, press,<br />

communitarian stable) with no current use or considered obsolete;<br />

• Stagnation, ab<strong>and</strong>onment <strong>and</strong> aging, more relevant after 2005. Building increased but houses<br />

are closed. Small medical centers <strong>and</strong> schools, local <strong>and</strong> regional services for farmers were<br />

disabled <strong>and</strong> relative recent infrastructures are closed <strong>and</strong> ab<strong>and</strong>oned. A great majority of the<br />

inhabitants is older than 60 <strong>and</strong> there is serious lack of children <strong>and</strong> young.<br />

Informants have also reported noticeable changes in l<strong>and</strong> use <strong>and</strong> cover such as no cattle grazing,<br />

quite ab<strong>and</strong>oned meadows <strong>and</strong> arable l<strong>and</strong>s, the absence of once usual crops, for instance, flax<br />

<strong>and</strong> hops, the afforestation of individual fields, more diverse homegardens adjacent to houses,<br />

the presence of many cultivated <strong>and</strong> exotic ornamentals, <strong>and</strong> fenced fields <strong>and</strong> plots.<br />

Some of these topics that were also identified <strong>and</strong> perceived as probable causes of l<strong>and</strong>scape<br />

changing are detailed <strong>and</strong> summarized as follows.<br />

3.1 Cultural heritage <strong>and</strong> aesthetical values<br />

For a long time, villages’ subsistence was based mainly on forestry, pastoralism (cattle, sheep<br />

<strong>and</strong> goats) <strong>and</strong> sustainable farming systems with specific gardening techniques. People’s food<br />

<strong>and</strong> medical needs relied on materials found in the natural surroundings, on small-scale animal<br />

breeding, on fishing <strong>and</strong> hunting <strong>and</strong> on self-sufficient <strong>and</strong> subsistence oriented agriculture.<br />

Wild-gathered species were an important supplement <strong>and</strong> alternative to the regular diet <strong>and</strong><br />

often used to prepare homemade remedies for primary healthcare <strong>and</strong> treatment of human <strong>and</strong><br />

animal diseases. Arable crops, scrubl<strong>and</strong> <strong>and</strong> woods provided food <strong>and</strong> supplied other basic<br />

needs, such as fuel, domestic tools, textiles <strong>and</strong> building raw materials. At times, surpluses of<br />

grains, chestnuts, potatoes, livestock, textiles, h<strong>and</strong>icrafts, charcoal <strong>and</strong> wood were traded or<br />

sold, to generate extra income. Mining, smuggling <strong>and</strong> other men activities complemented the<br />

household income.<br />

According to informants’ testimony, those were days of great activity in the villages; women<br />

<strong>and</strong> children were active plant gatherers <strong>and</strong> foragers, while most of the men cultivated field<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Carvalho et al. 2010. Connecting l<strong>and</strong>scape conservation <strong>and</strong> management with traditional ecological knowledge<br />

477<br />

crops, worked in the woods or had jobs outside the community in farming, mining, road works<br />

<strong>and</strong> reforestation programs. Over time, this close relationship between people <strong>and</strong> their natural<br />

<strong>and</strong> agricultural environment has led to the development of a rich knowledge base on plants,<br />

plant uses <strong>and</strong> related practices (Carvalho <strong>and</strong> Morales 2010).<br />

Key informants have emphasized that over the recent two to five decades people’s adaptive<br />

management of natural resources has built a multifunctional, productive <strong>and</strong> diverse l<strong>and</strong>scape.<br />

L<strong>and</strong>-use system respected a circular configuration with settlements in the middle; surrounding<br />

the houses, homegardens, arable l<strong>and</strong>s, scrubl<strong>and</strong>, woods <strong>and</strong> crop rotation (rye - more or less<br />

long fallow) following a decreasing gradient of soil fertility but increasingly slope <strong>and</strong> distance<br />

to center; meadows were <strong>and</strong> still are transversal to theses aureoles (Aguiar et al. 2009).<br />

This traditional l<strong>and</strong>scape is considered part of their cultural heritage <strong>and</strong> has embedded<br />

intangible values such as dwelling <strong>and</strong> aesthetical values, local tradition, neighborly <strong>and</strong> intergenerational<br />

relations.<br />

3.2 Disabling traditional agricultural activities<br />

Cereal production, crop rotation <strong>and</strong> animal husb<strong>and</strong>ry are locally considered as linked practices<br />

<strong>and</strong> skills that may not last alone because they are meaningless without each other. As they<br />

explained growing wheat or rye usually provided three sub-products: straw for litter <strong>and</strong> basket<br />

weaving, grain for selling in the market <strong>and</strong> to keep at home, stubbles <strong>and</strong> fallows for feeding<br />

sheep in late summer <strong>and</strong> after the first autumn rains.<br />

People were able to remember the enlargement of the areas assigned to rye, wheat <strong>and</strong> fodder<br />

production during the 1950s, as well as, the satisfactory performances of local varieties well<br />

adapted <strong>and</strong> the consequent increase of cattle <strong>and</strong> sheep, which in turn, also concerned the<br />

management of the scrubl<strong>and</strong>s, meadows <strong>and</strong> pastures, fallows <strong>and</strong> stubbles. Cycles of slash<strong>and</strong>-burn,<br />

cultivation, <strong>and</strong> scrub were still common in the 1970s. Species from the scrubl<strong>and</strong><br />

were used as fertilizer, litter, pasture, firewood, <strong>and</strong> some to make charcoal.<br />

Considering the informants reports, there is a general idea that traditional agricultural farming<br />

systems have been affected by a succession of CAP (Common Agricultural Policy) reforms <strong>and</strong><br />

‘flanking’ measures, beginning in 1992 <strong>and</strong> continuing for twelve years, that cancelled or<br />

reduced some crop subsidies, introduced new varieties <strong>and</strong> imposed strict production conditions.<br />

These measures disappointed many local farmers <strong>and</strong> caused the ab<strong>and</strong>onment of several crops<br />

(such as cereal grain, potatoes, fodder, fibers <strong>and</strong> hop). Breeders experienced some difficulties<br />

to meet municipal ordinances <strong>and</strong> innovative requirements concerning animal welfare <strong>and</strong><br />

veterinary care, which require continued technical assistance. New policies constrained the<br />

ability of small farmers to diversify <strong>and</strong> reduced the mosaic of farming activity. Agriculture was<br />

suddenly viewed as an impossible task without competitive advantages because of rising<br />

production costs versus low profits <strong>and</strong> uncertain wages.<br />

Perceived main indicators of l<strong>and</strong>scaping changing due to non prevailing agricultural practices<br />

are ab<strong>and</strong>oned arable l<strong>and</strong>s <strong>and</strong> meadows that progressively exhibit a different floristic<br />

composition <strong>and</strong> scrubl<strong>and</strong> represented by a tallest stratum with increased risk of wildfires.<br />

As meadows are not cut for hay or grazed the early colonizers will be shaded out when woody<br />

plants become well-established. Several medicinal plants often gathered in these fields are no<br />

more available.<br />

3.3 Perennial rather than seasonal crops<br />

In general, afforestation of farml<strong>and</strong> is regard as a good alternative to seasonal crops <strong>and</strong><br />

ab<strong>and</strong>onment because it allows absenteeism, provides income <strong>and</strong> represents a patrimony for<br />

future generations.<br />

Chestnut, walnut tree, cherry tree <strong>and</strong> red oak are the most mentioned species for afforestation.<br />

Arable l<strong>and</strong>s, dry prairies <strong>and</strong> common l<strong>and</strong>s have been afforested for both timber <strong>and</strong> fruit<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Carvalho et al. 2010. Connecting l<strong>and</strong>scape conservation <strong>and</strong> management with traditional ecological knowledge<br />

478<br />

production. In wet meadows, fast-growing hybrid poplars are grown on plantations <strong>and</strong> sold for<br />

pulpwood <strong>and</strong> as inexpensive hardwood timber, used for pallets <strong>and</strong> cheap plywood.<br />

Altough it seems a good alternative, several informants comented that there is a risk of plant<br />

diseases, especially ink-desease in chestnut. Moreover, seasonal labor for fruit recollection <strong>and</strong><br />

species management is considered expensive, scarce <strong>and</strong> difficult to hire.<br />

Afforestation changes the traditional lanscape mosaic, as there are scattered afforested patches,<br />

combined with annual crops, meadows <strong>and</strong> scrubl<strong>and</strong>.<br />

3.4 New food plants <strong>and</strong> new herbaceous <strong>and</strong> woody ornamentals<br />

Floral composition of homegardens <strong>and</strong> new green spaces inside the settlements are also signals<br />

of trasnformation in l<strong>and</strong> use <strong>and</strong> traditional l<strong>and</strong>scape. According to female informants, the<br />

number of cultivated species has increased with the introduction of a wide range of greens <strong>and</strong><br />

ornamental species in the last three decades. These plants or propagation materials have been<br />

brought from remote areas, exchanged between relatives <strong>and</strong> neighbours or bought from<br />

retailers at the local markets. For instance, the gathering <strong>and</strong> consumption of wild edible plants<br />

is in steady decline throughout the area; therefore women have brought some of the most<br />

popular plants used as food additives <strong>and</strong> beverages from the wild to grow in their homegardens,<br />

in order to make them easily available.<br />

Both a decline of agriculture <strong>and</strong> recent demographic trends have generated new approaches to<br />

homegardens. In former times they were less diverse because other agricultural activities such<br />

as forestry, grain production <strong>and</strong> animal husb<strong>and</strong>ry were considered much more important for<br />

the household economy. Food production in homegardens was very limited <strong>and</strong> they were<br />

mainly used to grow fodder <strong>and</strong> flax to make linen.<br />

In order to replicate urban lifestyles, villages’authorities created new areas <strong>and</strong> gardens where<br />

they have introduced exotic herbaceous <strong>and</strong> woody ornamentals which are, whenever possible,<br />

quickly propagated <strong>and</strong> used in homegardens. These ornamentals are also use in rituals <strong>and</strong><br />

cerimonies <strong>and</strong> have taken the place of wild species previously harvested from the forest by<br />

women <strong>and</strong> children.<br />

More diverse homegardens <strong>and</strong> new ornamental gardens are also perceived as new structural<br />

components of l<strong>and</strong>scapes<br />

4. Discussion<br />

Along the interviews, new farming practices, ab<strong>and</strong>onment of farming <strong>and</strong> husb<strong>and</strong>ry activities,<br />

a better mobility <strong>and</strong> a new concept of residential housing were the most mentioned causes for<br />

l<strong>and</strong>scape changing. Key informants perceive that young <strong>and</strong> some middle aged people value<br />

some of these changes, which they consider less hard-working <strong>and</strong> a symbol of modernity<br />

allowing a more like urban lifestyle (e.g. weekends <strong>and</strong> holidays). Others regret actual<br />

l<strong>and</strong>scape transformations which they view as a signal of ab<strong>and</strong>onment, waste of resources,<br />

reprehensibly unproductive. Nevertheless, most of the informants are aware of a dynamic<br />

process that is taking place <strong>and</strong> conscious that l<strong>and</strong>scape, like themselves, must adapt to<br />

changing times.<br />

Beginning as children, some people have learned how to discover <strong>and</strong> underst<strong>and</strong> the signs of<br />

nature <strong>and</strong> to observe changes in the l<strong>and</strong>scape. However, they have also shaped l<strong>and</strong>scape<br />

according to their own beliefs <strong>and</strong> material needs. This adaptive knowledge is often a practical<br />

one, based on empirical observation <strong>and</strong> long experience, <strong>and</strong> transmitted through oral traditions.<br />

Such knowledge is not merely of academic or historical interest but is fundamental to<br />

maintaining cultural continuity <strong>and</strong> identity <strong>and</strong>, possibly, could play a role in achieving<br />

sustainable use of plant resources in the future. It is also useful for providing more realistic<br />

evaluations of environment, natural resources <strong>and</strong> production systems. TEK may improve<br />

success by involving local people in the planning processes. Therefore TEK <strong>and</strong> local<br />

conceptions can be considered important tools for l<strong>and</strong>scape conservation <strong>and</strong> management.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Carvalho et al. 2010. Connecting l<strong>and</strong>scape conservation <strong>and</strong> management with traditional ecological knowledge<br />

479<br />

By interviewing specifically on folk nomenclature <strong>and</strong> identification of useful plants we<br />

observed that the loss of TEK <strong>and</strong> loss of vocabulary begin with people aged less than 50 years.<br />

The lost of traditional knowledge <strong>and</strong> local categorization <strong>and</strong> naming are not completely<br />

coupled: a few interviewees of the middle generation seem to be often able to remember the<br />

names of plants, but not to identify them or to explain their traditional use or to find the sites<br />

were these plants were usually gathered.<br />

It became clear that in the past thirty years, homegardens have become areas of in situ <strong>and</strong> ex<br />

situ conservation for both nostalgic <strong>and</strong> pragmatic reasons. Some crops <strong>and</strong> l<strong>and</strong>races are no<br />

longer cultivated in arable fields <strong>and</strong> wild species are threatened by new access roads, wild fires,<br />

<strong>and</strong> reforestation activities.<br />

Although some of the components may stay unchanged, much of aesthetic, historical or<br />

cultural value of rural l<strong>and</strong>scapes remains to be inventoried <strong>and</strong> recorded which is<br />

urgent before it disappears.<br />

References<br />

Aguiar, C., Rodrigues, O., Azevedo, J. <strong>and</strong> Domingos, T., 2009. Montanha. In: H. M. Pereira,<br />

T. Domingos, L. Vicente, <strong>and</strong> V. Proença (Eds.). Ecossistemas e bem estar humano.<br />

Avaliação para Portugal do Millennium Ecosystem Assessment. Lisboa: <strong>Escola</strong>r Editora.<br />

Alexiades, M. N., 1996. Collecting Ethnobotanical Data: an Introduction to Basic Concepts <strong>and</strong><br />

Techniques. In:M. N. Alexiades (Ed.). Selected Guidelines for Ethnobotanical Research:<br />

a Field Manual. New York: The New York Botanical Garden: 53-94.<br />

Antrop, M., 2005. Why l<strong>and</strong>scapes of the past are important for the future. L<strong>and</strong>scape <strong>and</strong><br />

Urban Planning, 70: 21–34.<br />

Calvo-Iglesias, M. S., Fra-Paleo, U. <strong>and</strong> Diaz-Varela, R. A., 2009. Changins in farming system<br />

<strong>and</strong> population as drivers of l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong>scape dynamics: the case of enclosed <strong>and</strong><br />

semi-openfield systems in Northern Galicia (Spain). L<strong>and</strong>scape <strong>and</strong> Urban Planning, 90:<br />

168-177.<br />

Carvalho, A. M. <strong>and</strong> Morales, R., 2010. Persistence of Wild Food <strong>and</strong> Wild Medicinal Plant<br />

Knowledge in a North-Eastern Region of Portugal. In: M. Pardo de Santayana, A.<br />

Pieroni, <strong>and</strong> R. Puri (Eds.). Ethnobotany in the New Europe: People, Health <strong>and</strong> Wild<br />

Plant Resources. Oxford, UK: Berghahn Books.<br />

Carvalho, A. M., 2010. Plantas y sabiduría popular del Parque Natural de Montesinho. Un<br />

estudio etnobotánico en Portugal. Biblioteca de Ciencias, Consejo <strong>Superior</strong> de<br />

Investigaciones Científicas, Madrid, 450 p.<br />

Council of Europe, 2000. European L<strong>and</strong>scape Convention. ETS No. 176, Publishing Division,<br />

Strasbourg.<br />

Cunningham, A. B., 2001. Applied ethnobotany. People, wild plant use <strong>and</strong> conservation.<br />

Earthscan Publications, London, 300 p.<br />

Frazão-Moreira, A., Carvalho, A. M. <strong>and</strong> Martins, M. E., 2009. Local ecological knowledge<br />

also ‘comes from books’: cultural change, l<strong>and</strong>scape transformation <strong>and</strong> conservation of<br />

biodiversity in two protected areas in Portugal. Anthropological Notebooks 15 (1): 27–36.<br />

Martin, G., 1995. Ethnobotany: a methods manual. Earthscan Publications, London, 268 p.<br />

Plieninger, T., Hochtl, F. <strong>and</strong> Spek, T., 2006. Traditional l<strong>and</strong>-use <strong>and</strong> nature conservation in<br />

European rural l<strong>and</strong>scapes. Environmental science & pol icy, 9: 317-321.<br />

Ramos, M. T., 2008. Património vegetal e etnobotãnico no Planalto Mir<strong>and</strong>ês. Master degree<br />

Thesis in Gestão e Conservação da Natureza, Universidade dos Açores & Instituto<br />

Politécnico de Bragança, Portugal, 80p.<br />

Vos, M. <strong>and</strong> Meekes H., 1999. Trends in European cultural l<strong>and</strong>scape development:<br />

perspectives for a sustainable future. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 46: 3-14.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Deconchat et al. 2010. Identification <strong>and</strong> characterization of forest edge segments for mapping edge diversity<br />

480<br />

Identification <strong>and</strong> characterization of forest edge segments for<br />

mapping edge diversity in rural l<strong>and</strong>scapes.<br />

Marc Deconchat, Audrey Alignier, Philippe Espy & Sylvie Ladet<br />

UMR1201, Dynafor, INRA, INPT, F-31326 Castanet tolosan, France<br />

Abstract<br />

<strong>Forest</strong> edges are key components of rural l<strong>and</strong>scapes <strong>and</strong> they influence major ecological<br />

processes. There is a variability of forest edges, according to their physiognomy, orientation,<br />

history <strong>and</strong> topography, but few methods are available to identify automatically these different<br />

types of edges <strong>and</strong> to map them at a large scale. We propose to identify edge segments based on<br />

morphological subdivision of forest boundaries. These segments can be mapped <strong>and</strong><br />

characterized by other spatial data. A procedure based on Arcgis tools, applied on the output<br />

from GUIDOS tools, is used to identify edge segments. We provide examples of edge segments<br />

descriptors obtained from a digital elevation model <strong>and</strong> from a l<strong>and</strong>cover map. Results showed<br />

the feasibility of the automatic mapping of edge variability over a large scale. It opens new<br />

perspectives for the analysis of l<strong>and</strong>scape dynamics <strong>and</strong> their effects on biodiversity.<br />

Keywords: <strong>Forest</strong> edge, GIS, edge segment, morphology<br />

1. Introduction<br />

<strong>Forest</strong> edges are key components of rural l<strong>and</strong>scapes by their influence on major ecological<br />

processes important for biodiversity conservation (Murcia, 1995). A better underst<strong>and</strong>ing <strong>and</strong><br />

measurement of the extent <strong>and</strong> of the variability of edge effects in rural l<strong>and</strong>scapes is required<br />

by l<strong>and</strong> managers to base their decisions on accurate estimations.<br />

<strong>Forest</strong> edge influence on both microenvironmental conditions <strong>and</strong> vegetation depends on edge<br />

type (Alignier & Deconchat, 2010). There is a variability of forest edges, according to their<br />

physiognomy, orientation, history <strong>and</strong> topography, but few methods are available to identify<br />

automatically these different types of edges <strong>and</strong> to map them at a large scale. Essen et al (2006)<br />

proposed a method based on aerial photography interpretation with a systematic sampling of<br />

edges by a square grid. This method suffers from the difficulty to be applied on very large areas,<br />

the bias associated to the size of the sampling grid <strong>and</strong> the impossibility to produce a synthetic<br />

map of edge diversity. GUIDOS is a GIS tool developed for the European Union in order to<br />

measure <strong>and</strong> map habitat fragmentation (Vogt et al., 2007). In this method, edges are defined as<br />

the contour of large tracks of forest, they can be mapped but there is no evaluation of their<br />

diversity. Fragstat (McGarigal et al., 2002) is a set of GIS tools aiming at measuring<br />

fragmentation by patch metrics. It includes several outputs describing edge characteristics, but<br />

with few possibilities to take into account their diversity. Zeng & Wu (2005) introduced the<br />

concept of edge segment as the basic component for computing edge-based metrics to describe<br />

l<strong>and</strong>scape fragmentation. This approach open up new perspective <strong>and</strong> define the basis for a<br />

systematic <strong>and</strong> coherent method to map edge diversity in rural l<strong>and</strong>scape.<br />

The aim of the presentation is to introduce the details of a GIS method to identify <strong>and</strong><br />

characterize edge segments in l<strong>and</strong>scape, defined from morphological subdivision of forest<br />

boundaries, in the same general framework proposed by GUIDOS. This method, called<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Deconchat et al. 2010. Identification <strong>and</strong> characterization of forest edge segments for mapping edge diversity<br />

481<br />

Mapedge, is a step toward l<strong>and</strong>scape metrics that put more emphasis on interactions between<br />

l<strong>and</strong> covers than classical area-based metrics (McGarigal et al., 2009).<br />

2. Methodology<br />

The central idea of the method is to subdivide the contours of the forest patches in a set of<br />

segments that can be characterized by them or by GIS queries on other data maps (i. e. digital<br />

elevation model <strong>and</strong> l<strong>and</strong> cover map).<br />

Segmentation of edges<br />

There are several ways to cut forest contours in a set of edge segments. We chose to base our<br />

segmentation on forest edge morphology. We followed on this point the position adopted by<br />

GUIDOS authors who define different parts of forest cover according to morphological<br />

parameters (area, shape index, etc.) (Vogt et al., 2007). With this option, our method is forestcentered<br />

<strong>and</strong> edge segments are defined only by forest characteristics <strong>and</strong> not by elements<br />

outside the forest. Each step needs to select one or several parameters that may change the final<br />

result.<br />

In a first step (Figure 1), the input of Mapedge is a forest/non-forest cover map, obtained in our<br />

case from a remote sensed image (SPOT) classified with supervised control. In a second step<br />

(Figure 1), this map is processed through a GUIDOS tool, MPSA (for “Morphological Spatial<br />

Pattern Analysis”) in order to identify true forest patches <strong>and</strong> to discard the other smaller<br />

forested l<strong>and</strong>scape elements (i. e. hedgerows, branches) for which edge segmentation is not<br />

relevant. By convenience, we selected a depth of edge of 20 m (2 pixels) (Table 1). The output<br />

of MSPA was a classified forest map from which forest edges <strong>and</strong> cores were extracted.<br />

After that, the method is based on Arcgis 9.3 tools <strong>and</strong> several additional extensions<br />

(ETGeowizards 9.7; Hawth’s analysis tools 3.27; EasyCalculate 5.0) or free download scripts<br />

(Vba <strong>and</strong> Python languages). In a near future, all the processing operations will be encapsulated<br />

in a single application turned under Model Builder.<br />

The contours of the forest patches are then vectorized <strong>and</strong> generalized in order to obtain a<br />

smooth shape of the patches (Figure 1). The parameter of this step is T (Tolerance for<br />

generalization) (Table 1). The forest patch polygons are converted to polylines. Polylines are<br />

then split in segments <strong>and</strong> each segment is numbered <strong>and</strong> oriented, allowing to know on which<br />

side the forest is. Orientation, length, etc. of the segments can be extracted easily from the map<br />

<strong>and</strong> tabulated in order to provide edge-based metrics of the l<strong>and</strong>scape structure. An attribute<br />

table is attached to the map of edge segments <strong>and</strong> contains a set of variables describing each<br />

segment. This table can be filled by data obtained from map queries <strong>and</strong> introduced in statistical<br />

analysis of l<strong>and</strong>scape structure.<br />

Map queries<br />

We decided that the query of other maps will be based on transects perpendicular to the medium<br />

point of the edge segments towards the open habitat. We chose a fixed length of transect based<br />

on depth of edge (2*DE)(Table 1).<br />

As topography is known to be of prime importance for ecological edge effects, we combined the<br />

map of edge segments with a Digital Elevation Model (DEM). According to the difference of<br />

elevation along the transect <strong>and</strong> along the edge segment, we measured the position of the edge<br />

relatively to the main slope. We were then able to identify 3 classes: 1) edge where the open<br />

habitat is at a higher elevation than the edge (downslope), 2) edge where the open habitat is at a<br />

lower elevation than the edge (upslope) <strong>and</strong> 3) the other cases where the slope of the edge<br />

segment is higher than the slope of the transect (inslope).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Deconchat et al. 2010. Identification <strong>and</strong> characterization of forest edge segments for mapping edge diversity<br />

482<br />

Adjacent l<strong>and</strong> cover was measured with a spatial query between the summit of the transect <strong>and</strong> a<br />

l<strong>and</strong> cover map obtained in our case from the same remote sensed image treatment as the input<br />

forest map.<br />

Legend<br />

In order to be able to map the diversity of edge segment, for visual interpretations <strong>and</strong><br />

communication, we defined a system of symbols for edge segments based on the transects. They<br />

appeared on the map as pin-like symbols, with a color <strong>and</strong> a shape of the pin-head defined<br />

according to the characteristics of the edge segment (Figure 1).<br />

3. Results<br />

Figure 2 is an example of a result map with a focus on small area in order to show clearly the<br />

symbols that were used to display the different types of edge segments. Visual interpretation of<br />

the map, which would be statistically confirmed by an analysis of the attribute table, indicates<br />

that, in this very small area, meadow seemed to be the more frequent as an adjacent l<strong>and</strong> cover<br />

for woods 3 <strong>and</strong> 4 than for the others, more in contact with crops. Wooded elements of the<br />

l<strong>and</strong>scape (loop, branch, forest defined according to GUIDOS) were in the near vicinity of all<br />

the woods.<br />

4. Discussion<br />

Interfaces <strong>and</strong> interactions between forests <strong>and</strong> the other l<strong>and</strong> covers in l<strong>and</strong>scape are becoming<br />

focus questions for l<strong>and</strong> management in the perspective of global changes. Our method,<br />

Mapedge, contributes to provide managers with a new set of tools to take these interactions into<br />

account in their decisions. For example, metrics based on edges can provide measures of<br />

l<strong>and</strong>scape fragmentation to be compared with area based metrics (McGarigal et al., 2009). Some<br />

edge types may be more sensitive to l<strong>and</strong>scape changes than other (Taylor et al., 2008). The<br />

feasibility of edge diversity mapping has been demonstrated with our first results.<br />

Mapedge is an efficient method to identify <strong>and</strong> characterize edge segments in l<strong>and</strong>scape.<br />

Despite the need to combine several tools to process the data, the Mapedge methodology is<br />

rather simple <strong>and</strong> easy to apply. The number of parameters has been reduced as much as<br />

possible <strong>and</strong> rules have been defined to select their values. We are now testing several sets of<br />

parameters in order to assess how they influence the final output of the method (sensibility<br />

analysis). In a near future, we will set up a unique tool that will apply all the treatments which<br />

are separated for the moment. From a l<strong>and</strong>scape ecology point of view, our method will provide<br />

new l<strong>and</strong>scape descriptors that will give a higher role to interfaces in l<strong>and</strong>scape analysis.<br />

References<br />

Alignier, A., Deconchat, M., 2010. Variability of forest edge effect on vegetation implies to<br />

reconsider its assumed hypothetical pattern. Applied Vegetation Science (submitted).<br />

Esseen, P.A., Jansson, K.U. <strong>and</strong> Nilsson, M., 2006. <strong>Forest</strong> edge quantification by line intersect<br />

sampling in aerial photographs. <strong>Forest</strong> Ecology <strong>and</strong> Management, 230:32-42.<br />

McGarigal, K., S. A. Cushman, M. C. Neel, <strong>and</strong> E. Ene. 2002. FRAGSTATS: Spatial Pattern<br />

Analysis Program for Categorical Maps. Computer software program produced by the<br />

authors at the University of Massachusetts, Amherst. Available at the following web site:<br />

www.umass.edu/l<strong>and</strong>eco/research/fragstats/fragstats.html<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Deconchat et al. 2010. Identification <strong>and</strong> characterization of forest edge segments for mapping edge diversity<br />

483<br />

McGarigal, K., Tagil, S. <strong>and</strong> Cushman, S., 2009. Surface metrics: an alternative to patch metrics<br />

for the quantification of l<strong>and</strong>scape structure. L<strong>and</strong>sc. Ecol., 24:433-450.<br />

Murcia, C., 1995. Edge effects in fragmented forests - implications for conservation. Trends in<br />

Ecology & Evolution, 10:58-62.<br />

Taylor, R.S., Oldl<strong>and</strong>, J.M. <strong>and</strong> Clarke, M.F., 2008. Edge geometry influences patch-level<br />

habitat use by an edge specialist in south-eastern Australia. L<strong>and</strong>sc. Ecol., 23:377-389.<br />

Vogt, P., Riitters, K.H., Estreguil, C., Kozak, J. <strong>and</strong> Wade, T.G., 2007. Mapping spatial patterns<br />

with morphological image processing. L<strong>and</strong>sc. Ecol., 22:171-177.<br />

Zeng, H. <strong>and</strong> Wu, X.B., 2005. Utilities of edge-based metrics for studying l<strong>and</strong>scape<br />

fragmentation. Computers, Environment <strong>and</strong> Urban Systems, 29:159-178.<br />

Table 1: Description of the paramaters used in the different step of the process of Mapedge.<br />

Step Process Parameter Value<br />

1 Classification Resolution of SPOT 5 10 m<br />

Foreground Connectivity, for a set of 3 x 3 pixels the<br />

center pixel is connected to its adjacent neighboor pixels<br />

by having either.<br />

Connectivity<br />

= 8 pixels<br />

2<br />

Treatment under<br />

Guidos :<br />

Morphological Spatial<br />

Pattern Analysis<br />

3 Generalization<br />

5<br />

Creation of transect 90°<br />

Calculation<br />

L<strong>and</strong><br />

cover<br />

Edge Depth.<br />

Transition pixels are those pixels of an edge or a<br />

perforation where the core area intersects with a loop or<br />

a bridge. If Transition is set to 0 then the perforation <strong>and</strong><br />

the edges will be closed core boundaries.<br />

Intext allows distinguishing internal from external<br />

features, where internal features are defined as being<br />

enclosed by a Perforation. The default is to enable this<br />

distinction which will add a second layer of classes to the<br />

seven basic classes.<br />

Generalize tolerance reduces the number of vertices<br />

required to represent a polygon, the features of a polygon<br />

layer using the Douglas-Poiker algorithm.<br />

Script python what create perpendicular lines, takes a<br />

line shapefile <strong>and</strong> generates perpendicular lines to each<br />

record with length expected.<br />

Intersects the end-point of transect to l<strong>and</strong> cover at a<br />

point distance expected.<br />

Get Z characteristics with Digital Elevation Model<br />

Compare the edge to the transect segment with same<br />

Slope of length<br />

the edge Result :<br />

- Inslope with tolerance<br />

- Downslope<br />

- Upslope<br />

Cardinal<br />

Calculates the angle of a polyline in degrees<br />

orientation<br />

Size = 2<br />

pixels<br />

Transition =<br />

0<br />

Intext = 0<br />

20 meters<br />

40 meters<br />

40 meters<br />

25 meters<br />

resolution<br />

40 meter<br />

Tolerance 1<br />

meter<br />

360 degrees<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Deconchat et al. 2010. Identification <strong>and</strong> characterization of forest edge segments for mapping edge diversity<br />

484<br />

Figure 1 : GIS method in Arcgis using default functionalities of ArcToolbox <strong>and</strong> additional tools<br />

(extensions <strong>and</strong> scripts). In part A, we give the most important steps of the Mapedge method on small<br />

image composed with 2 woodlots. In step 1, input data come from supervised classification of satellite<br />

images (SPOT 5). In step 2, we used GUIDOS tool to make MPSA (Morphological Spatial Pattern<br />

Analysis). In step 3, we combine the 2 raster layers <strong>and</strong> convert to vector data <strong>and</strong> generalize to simplify<br />

features. In step4, we check direction of vertices <strong>and</strong> label them. In step 5, we create transect features<br />

from rotation of vertices <strong>and</strong> use them to calculate edge descriptors (about interface with adjacent l<strong>and</strong><br />

cover; about topography with Digital Elevation Model <strong>and</strong> orientation). We will apply the same method<br />

at l<strong>and</strong>scape level (several hundred polygons of woodlots) <strong>and</strong> transfer output results to statistical<br />

analyses. In Part B, we give legend of symbols.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Deconchat et al. 2010. Identification <strong>and</strong> characterization of forest edge segments for mapping edge diversity<br />

485<br />

Figure 2: Overview map of all descriptors calculated for test image made with 4 woodlots. We obtain 34<br />

edge segments. We symbolize cardinal orientation on edge segment, slope of edge on<br />

perpendicular transect <strong>and</strong> l<strong>and</strong> cover on pin-head of the transect.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Gaspar et al. 2010. Visibility analysis <strong>and</strong> visual diversity assessment in rural l<strong>and</strong>scapes<br />

486<br />

Visibility analysis <strong>and</strong> visual diversity assessment in rural l<strong>and</strong>scapes<br />

José Gaspar 1* , Beatriz Fidalgo 1 , David Miller 2 , Luis Pinto 3 & Raúl Salas 1<br />

1<br />

Instituto Politécnico de Coimbra, <strong>Escola</strong> <strong>Superior</strong> Agrária, Coimbra, Portugal<br />

2<br />

Macaulay L<strong>and</strong> Use Research Institute, Aberdeen, Scotl<strong>and</strong>, United Kingdom<br />

3<br />

Independent Consultant, Aveiro, Portugal<br />

Abstract<br />

Analysing the visibility of l<strong>and</strong>scape features is usually based on methods which calculate the<br />

number of locations from which such features are visible. However, visibility analysis using<br />

Geographic Information Systems (GIS) suffers from several limitations which can significantly<br />

influence the results. In this study the authors have tested the differences in visibility through<br />

time, <strong>and</strong> used a Digital Elevation Model which includes the heights of l<strong>and</strong> cover features.<br />

The results provided information about the visibility of each l<strong>and</strong> cover type, at each of three<br />

dates: 1954, 1974 <strong>and</strong> 1995. The derived maps of visibility were used as inputs to analysis of<br />

the visual content of the l<strong>and</strong>scape, which were then used to evaluate the visual diversity at each<br />

date, <strong>and</strong> to provide an input to the derivation of information on l<strong>and</strong>scape visual quality. The<br />

results show significant differences in both visual content <strong>and</strong> diversity through time, <strong>and</strong><br />

illustrate a difference between the potential perceptions of change, <strong>and</strong> the real extent of change<br />

in an area.<br />

Keywords: visibility analysis, l<strong>and</strong>scape visual diversity, l<strong>and</strong>scape content<br />

1. Introduction<br />

The planning <strong>and</strong> management of l<strong>and</strong> use requires an underst<strong>and</strong>ing of the benefits <strong>and</strong> impacts<br />

of change with respect to environmental, economic <strong>and</strong> social objectives. One aspect of<br />

management which links across these three broad headings is in relation to the visual l<strong>and</strong>scape,<br />

<strong>and</strong> the contributions made by the extent <strong>and</strong> distribution of different types of l<strong>and</strong> cover, <strong>and</strong><br />

associated uses. An analysis of diversity in the visual l<strong>and</strong>scape, <strong>and</strong> how it changes through<br />

time (Antrop, 2005), is an aid to underst<strong>and</strong>ing the wider significance of the impacts of drivers<br />

of ch’ange (Bell, 2001) (e.g. climate change; demographic change) <strong>and</strong> the consequent changes<br />

in l<strong>and</strong> cover (e.g. forest fire; urbanisation) than only the areas they occupy.<br />

An analysis of the visibility of l<strong>and</strong> cover types, <strong>and</strong> associated uses was undertaken, based on<br />

the methodology applied by Miller (2001), Gaspar et al. (2002) <strong>and</strong> Ode (2003). The outputs<br />

provided a basis for deriving indicators of visual diversity in the l<strong>and</strong>scape, <strong>and</strong> the visual<br />

complexity <strong>and</strong> content in the view.<br />

* Corresponding author. Tel.:+351 239802940 - Fax:+351 239802979<br />

Email address: jgaspar@esac.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Gaspar et al. 2010. Visibility analysis <strong>and</strong> visual diversity assessment in rural l<strong>and</strong>scapes<br />

487<br />

2. Methodology<br />

The methodology uses data on elevation, l<strong>and</strong> use <strong>and</strong> the height of features, focusing on forest<br />

areas, to derive <strong>and</strong> analyse the visibility of each l<strong>and</strong> use. The method is based on the<br />

measurment of the extent of l<strong>and</strong> from which different l<strong>and</strong> cover types may be visibile (Bolós,<br />

1992), on a cell-by-cell basis. The visibility calculations use a Digital Terrain Model (DTM), to<br />

which the height of the different forest species derived from growth models or inventory data<br />

were added, in order to obtain visibility results for each l<strong>and</strong> cover class across the municipality.<br />

Data were compiled for three dates: 1954, 1974 <strong>and</strong> 1995. The 1954 <strong>and</strong> 1974 data were<br />

digitised from paper maps of the l<strong>and</strong> use as mapped by the agriculture surveying services. The<br />

1995 data were produced by aerial photo-interpretation of false color imagery.<br />

The calculation of visibility is applied to the coverage of each type of l<strong>and</strong> cover, for each date,<br />

thus enabling an analysis of the combinations of visual influences of different types of l<strong>and</strong><br />

cover at any point in the l<strong>and</strong>scape (Miller, 2001; Gaspar <strong>and</strong> Fidalgo, 2002). Due to the extent<br />

of the study site, this study does not consider the effects of distance decay (Bolós, 1992) on the<br />

visibility of features (Gaspar et al., 2004).<br />

The individual visibility maps are combined to derive the l<strong>and</strong> cover types visible from each<br />

point in the municipality, as represented by a regular grid across the area. Three different levels<br />

of l<strong>and</strong>scape content were used to characterise the diversity of the views.<br />

3. Results<br />

To enable comparisons between l<strong>and</strong> use maps from different sources <strong>and</strong> with different<br />

classification schemes, a general classification scheme was developed with 9 classes (Gaspar,<br />

2005).<br />

Table 1. Extent of l<strong>and</strong> cover classes for 1954, 1974 <strong>and</strong> 1995.<br />

L<strong>and</strong> cover class<br />

1954 1974 1995<br />

Dry cropl<strong>and</strong> 13,7% 10,1% 8,5%<br />

Irrigated crops 7,5% 7,9% 2,7%<br />

Maritime pine 49,5% 64,4% 22,9%<br />

Eucalyptus 0,0% 1,7% 11,4%<br />

Broadleaved species 1,6% 1,4% 3,9%<br />

Other needle leaved species 0,1% 0,1% 0,2%<br />

Water 0,4% 0,4% 1,3%<br />

Other l<strong>and</strong> uses 26,7% 13,1% 47,5%<br />

Social areas 0,5% 0,9% 1,8%<br />

Total 100% 100% 100%<br />

The 1954 to 1974 period was marked by the progressive ab<strong>and</strong>onment of agriculture areas <strong>and</strong><br />

large plantations of Maritime pine (Pinus pinaster) (Table 1). Although tourism is now<br />

considered as one path towards a prosperous future, one of the main sources of income for the<br />

population is forest production, mainly associated to the paper industry, with the lower valleys<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Gaspar et al. 2010. Visibility analysis <strong>and</strong> visual diversity assessment in rural l<strong>and</strong>scapes<br />

488<br />

being occupied by Eucalyptus (Eucalyptus globulus) (Table 1). However, forest fires are a<br />

major driver of change of the l<strong>and</strong>scape.<br />

The results of the visibility analysis for each of the l<strong>and</strong> cover classes were combined to derive<br />

the values of visual diversity (Figure 1).<br />

Figure 1 L<strong>and</strong>scape visual diversity: number of l<strong>and</strong> cover types visible for 1954, 1974 <strong>and</strong><br />

1995.<br />

The changes in visual diversity through time shows a comparatively small change between 1954<br />

<strong>and</strong> 1974, but more significant changes by 1995, with an increase of the higher diversity values<br />

(7, 8 <strong>and</strong> 9), <strong>and</strong> a decrease of the lower diversity values. These changes will be influenced by<br />

the reduction in the overall number of patches of l<strong>and</strong> cover classes between 1954 <strong>and</strong> 1974, but<br />

an increase in the area <strong>and</strong> patches of Maritime pine <strong>and</strong> Other l<strong>and</strong> uses. The combination of<br />

these changes has led to an increase in the mean patch size, <strong>and</strong> the increase in the distance<br />

between patches of the same l<strong>and</strong> cover classes having a significant influence on the distribution<br />

of the visual diversity.<br />

The period between 1974 <strong>and</strong> 1995 shows a significant increase in higher visual diversity values,<br />

which can be explained by the increase of fragmentation, but also the new patches of Eucalyptus<br />

<strong>and</strong> Broadleaved species in highly visibile areas. The effect of forest fires also had a srong<br />

effect in terms of patch fragmentation, <strong>and</strong> provided conditions to increase the visibility of<br />

ab<strong>and</strong>oned agriculture areas.<br />

Table 2 L<strong>and</strong>scape content.<br />

L<strong>and</strong> cover class visibility<br />

1954 1974 1995<br />

30% 50% 70% 30% 50% 70% 30% 50% 70%<br />

Dry cropl<strong>and</strong> 1.0 1.2 0.6 0.3 0.6 0.5<br />

Irrigated crops<br />

Maritime pine 59.4 68.9 38.2 81.3 86.7 55.4 21.3 14.5 4.6<br />

Eucalyptus 0.6 0.8 9.2 6.9 1.5<br />

Broadleaved species 0.5 0.5<br />

Other conifers<br />

Water<br />

Other l<strong>and</strong> uses 14.2 17.8 10.7 4.2 7.0 3.3 41.1 40.7 26.8<br />

Social areas<br />

Maritime pine + Dry cropl<strong>and</strong> 1.9 1.4 0.3<br />

Maritime pine + Irrigated crops 0.5<br />

Maritime pine + Eucalyptus 1.1 4.6<br />

Other l<strong>and</strong> uses + Maritime pine 21.3 10.6 13.3<br />

Other l<strong>and</strong> uses + Eucalyptus 4.6<br />

Other l<strong>and</strong> uses + Broadleaved species 0.4<br />

Other l<strong>and</strong> uses + Water 0.3<br />

Total 98.3 87.9 49.6 99.3 94.8 58.8 96.2 63.1 32.9<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Gaspar et al. 2010. Visibility analysis <strong>and</strong> visual diversity assessment in rural l<strong>and</strong>scapes<br />

489<br />

The analysis of the changes in l<strong>and</strong>scape content for each cell through time was derived using<br />

the individual visibility maps (Table 2, <strong>and</strong> Figures 2, 3 <strong>and</strong> 4). Three levels of viewshed<br />

content were considered. The lower value considered was 30% of the cell visibility content, the<br />

intermediate value 50% <strong>and</strong> the highest 70%.<br />

The analysis of the results shows a significant dominance in the l<strong>and</strong>scape of the Maritime pine<br />

(1954 <strong>and</strong> 1974) (Table 2). For example, in 1974, this cover type was present in 55.4% of the<br />

area with a 70% level of l<strong>and</strong>scape content (Table 2). There is also an increase in the area where<br />

only one l<strong>and</strong> cover type dominates the l<strong>and</strong>scape view content.<br />

The l<strong>and</strong> cover change between 1974 <strong>and</strong> 1995 (Table 2) provided conditions for significant<br />

variations in terms of l<strong>and</strong>scape visual content. The l<strong>and</strong>scape is dominated by the Other l<strong>and</strong><br />

uses class in all levels of l<strong>and</strong>scape content (Figures 2, 3 <strong>and</strong> 4).<br />

The Maritime pine areas are the second largest class in terms of levels of visibility, which can<br />

be due to the presence in highly visible areas, <strong>and</strong> the significant increase in Eucalyptus in terms<br />

of both area <strong>and</strong> visibility.<br />

Figure 2 L<strong>and</strong>scape content 30% in 1995.<br />

The number of classes visible at the 30% level shows an increase in number <strong>and</strong> type when<br />

compared to the data for 1954 <strong>and</strong> 1974, which could be due to the increase of number of<br />

patches in each class, <strong>and</strong> related to the spatial distribution of these patches across the l<strong>and</strong>scape.<br />

Figure 3 L<strong>and</strong>scape content 50% in 1995.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Gaspar et al. 2010. Visibility analysis <strong>and</strong> visual diversity assessment in rural l<strong>and</strong>scapes<br />

490<br />

Figure 4 L<strong>and</strong>scape content 70% in 1995.<br />

4. Discussion<br />

This paper presents an approach to analysing visual diversity <strong>and</strong> content of l<strong>and</strong> cover <strong>and</strong><br />

change, using GIS tools. The l<strong>and</strong>scape diversity <strong>and</strong> content maps provide information for use<br />

in forest management <strong>and</strong> planning. The final analysis <strong>and</strong> evaluation of the results is ongoing,<br />

based on expert experience <strong>and</strong> opinion. This will include an assessment of their value in terms<br />

of parameters for use in planning tourist routes.<br />

References<br />

Antrop, M., 2005. Why l<strong>and</strong>scapes of the past are important for the future. L<strong>and</strong>scape <strong>and</strong><br />

Urban Planning, 70 (3-4), 21-34.<br />

Bell, S., 2001. L<strong>and</strong>scape pattern, perception <strong>and</strong> visualisation in the visual management of<br />

forests. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 54(1-4), 201-211.<br />

Bólos, M., 1992. Manual de ciencia del paisage - teoría, métodos y aplicaciones. Masson,<br />

Barcelona, 273 p.<br />

Fidalgo, B., 2005. Planeamento com análise multicritério em Paisagens Florestais. PhD<br />

Thesis in <strong>Forest</strong> Engineering, Universidade Técnica de Lisboa, Instituto <strong>Superior</strong> de<br />

Agronomia, Lisbon.<br />

Fidalgo, B., Gaspar, J., 2001. Utilização da fotointerpretação e indicadores cartográficos<br />

na caracterização do mosaico florestal à escala do município. IV Congresso Florestal<br />

Nacional (Évora 28-30 November 2001).<br />

Gaspar, J., 2005. A gestão da floresta e o planeamento do uso do solo, PhD Thesis in Applied<br />

Environmental Science, Universidade de Aveiro, Aveiro, 273 p.<br />

Gaspar, J., Fidalgo, B., Pinto, L., 2004. Visibilidade do uso do a diferentes distâncias – O<br />

contributo do projecto VisuL<strong>and</strong>s. ESIG 2004 (Lisbon 2-4 June 2004), 6 p.<br />

Gaspar, J, Fidalgo, B., 2002. Evolução do Uso do Solo e Avaliação do Valor Paisagístico e de<br />

Recreio na Área de Paisagem Protegida da Serra do Açor. Silva Lusitana, Vol. 10 (2),<br />

179-194.<br />

Gaspar, J., Miller, D., Fidalgo, B., 2001. Modelling the visual l<strong>and</strong>scape of protected areas. 2º<br />

Congresso Nacional da Conservação da Natureza (Lisbon October 2001).<br />

Miller, D., 2001. A method for estimating changes in the visibility of l<strong>and</strong> cover. L<strong>and</strong>scape<br />

<strong>and</strong> Urban Planning, Vol. 54, 91-104.<br />

Ode, Ǻ., 2003. Visual aspects in urban woodl<strong>and</strong> management <strong>and</strong> planning. Doctoral Thesis,<br />

Swedish University of Agricultural Sciences, Alnarp, Sweden, 110 p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes 2010. L<strong>and</strong>scape runoff, precipitation variation <strong>and</strong> reservoir limnology<br />

491<br />

L<strong>and</strong>scape runoff, precipitation variation <strong>and</strong> reservoir limnology<br />

A.M. Geraldes *<br />

CIMO, <strong>Escola</strong> <strong>Superior</strong> Agrária do Instituto Politécnico de Bragança Campus de<br />

Santa Apolónia, Bragança, Portugal<br />

Abstract<br />

L<strong>and</strong>scape runoff potential impact on reservoir limnology was indirectly evaluated by assessing<br />

the effect of precipitation variation on several water quality parameters, on Anabaena<br />

(Cyanophyta) <strong>and</strong> crustacean zooplankton abundances. The obtained results showed that total<br />

phosphorus increased with strong precipitation events whereas water transparency presented an<br />

opposite trend. Wet periods followed by long dry periods favored Anabaena dominance, which<br />

induced an accentuated decreasing on all crustacean zooplankton species abundance. Therefore,<br />

in a climate changing scenario these data are crucial to monitor <strong>and</strong> predict the effect of<br />

l<strong>and</strong>scape changes on aquatic ecosystem integrity <strong>and</strong> ultimately in water quality.<br />

Keywords:L<strong>and</strong>scape runoff impact,Precipitation,Reservoir limnology ,Water quality<br />

1. Introduction<br />

In Mediterranean region, both intensity <strong>and</strong> quantity of precipitation can vary markedly from<br />

one year to another. Such variability can generate different kinds of seasonal patterns,<br />

modifying water turnover time <strong>and</strong> changing the intensity of environmental <strong>and</strong> biological<br />

processes occurring in the water column of aquatic systems. Furthermore, the external loading<br />

of nutrients, organic matter <strong>and</strong> pollutants often increases with intensive precipitation events.<br />

The intensity <strong>and</strong> the magnitude of these loadings depend on l<strong>and</strong> use, vegetation cover <strong>and</strong><br />

l<strong>and</strong>scape patchiness. Azibo Reservoir is located in the Iberian Peninsula on the Portuguese part<br />

of international River Douro catchment. In this region, most precipitation occurs between<br />

October <strong>and</strong> March in a very irregular pattern from one year to another (Fig. 1). Total annual<br />

precipitation varies between is around 760 mm, <strong>and</strong> in a “normal” autumn/winter season total<br />

precipitation is around 500 mm (Instituto de Meteorologia, 2009). In contrast to what happens<br />

in other reservoirs in the region, water level fluctuations caused by human activity are not very<br />

accentuated in Azibo. Thus, this reservoir provides a good environment to study the potential<br />

effects of quantity <strong>and</strong> intensity of precipitation without the interference of internal disturbances<br />

generated by extreme anthropogenic water level fluctuations. The objective of the present<br />

research is to evaluate the potential impact of l<strong>and</strong>scape runoff on reservoir limnology.<br />

This was achieved indirectly through the assessment of the effects of precipitation<br />

variation on (1) environmental variables such as total phosphorous (TP), dissolved oxygen<br />

(DO), conductivity, pH, water transparency <strong>and</strong> chlorophyll a (Chl a); (2) Anabaena <strong>and</strong><br />

crustacean zooplankton species abundance.<br />

* Corresponding author. Tel.:351273303217 - Fax:351273331570<br />

Email address: geraldes@ipb.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes 2010. L<strong>and</strong>scape runoff, precipitation variation <strong>and</strong> reservoir limnology<br />

492<br />

2. Methodology<br />

Water, Anabaena <strong>and</strong> crustacean zooplankton samples were collected in the deepest point of the<br />

reservoir in four distinct hydrological years: (1) 2000/2001; (2) 2001/2002; (3) 2007/2008 <strong>and</strong><br />

(4) 2008/2009. Details on sampling methodology <strong>and</strong> laboratorial procedure are described in<br />

Geraldes <strong>and</strong> Boavida (2004; 2007). As most precipitation occurs between October <strong>and</strong> March<br />

samples collected during this period were classified as “winter season” the others were<br />

considered as “summer station”. A Kruskal-Wallis test was performed for each environmental<br />

variable <strong>and</strong> Anabaena abundance to determine whether mean values obtained for each year<br />

were significantly different. Non-metric Multidimentional Scaling (N-MSD) was used for<br />

expressing similarity between zooplankton samples. In this method samples are arranged in a<br />

continuum in such a way that those close together are similar <strong>and</strong> those which are far apart are<br />

dissimilar The statistical analysis were performed using SPSS 16 <strong>and</strong> CAP 3, respectively.<br />

3. Results<br />

The total precipitation recorded during the period of study is presented in Table 1. Significant<br />

inter-annual differences were found for TP (χ= 10.01; p= 0.001), for conductivity (χ= 34.71; p=<br />

0.000), for water transparency (χ= 19.74; p= 0.000) <strong>and</strong> for Anabaena densities (χ= 13.97; p=<br />

0.003). The maxima of TP <strong>and</strong> the minima of water transparency was recorded in the winter<br />

2000/2001. The low value of water transparency observed in winter 2001/2002 was a<br />

consequence of the increasing of Anabaena abundances (Table 1). The lowest densities for all<br />

crustacean zooplankton species was coincident with this period <strong>and</strong> with winter 2001/2002<br />

(Table 2). N-MSD for zooplankton samples also indicate the existence of seasonal <strong>and</strong> annual<br />

differences in crustacean zooplankton abundances (Figure 2).<br />

4. Discussion<br />

The observed variations in TP, water transparency <strong>and</strong> in Anabaena <strong>and</strong> crustacean zooplankton<br />

abundances were related to changes in rainfall/ l<strong>and</strong>scape runoff intensity. Similar results were<br />

obtained by authors studying reservoirs located in other regions influenced by Mediterranean<br />

climate (e.g. Armengol et al. 1999; Soria et al. 2000). The variation of l<strong>and</strong>scape runoff<br />

concomitantly with precipitation intensity was evidenced by the notice that TP concentrations in<br />

water samples obtained downstream of one of Azibo Reservoir tributary were 103μg l -1 at the<br />

beginning of the rainfall period decreasing subsequently to 76 μg l -1 <strong>and</strong> to 16 μg l -1 by the end<br />

of rainfall period (Geraldes, unpubl. data). Besides, the lowest mean values of TP <strong>and</strong> the<br />

highest values of water transparency noticed in the reservoir during the years with low<br />

precipitation reinforce the above evidence. The dominance of Anabeaena from October 2001 to<br />

December 2001 (dry winter) occurred subsequently to a wet winter. The nutrient increasing<br />

scenario created during the wet period (Winter 2000/01) followed by the environmental<br />

conditions (e.g. absence of water turbulence, larger water residence time <strong>and</strong> higher irradiance)<br />

created by the subsequent dry periods (summer 2001 <strong>and</strong> winter 2001/2002), had lead to<br />

Anabaena dominance over the other groups of phytoplankton assemblage. As this alga is not<br />

edible by the most of the species of crustacean zooplankton their densities decreased accentually.<br />

<strong>Change</strong>s in limnological parameters are related to variations in precipitation intensity, which<br />

ultimately influence l<strong>and</strong>scape runoff. Therefore, long term data series on reservoir abiotic <strong>and</strong><br />

biotic components will allow to underst<strong>and</strong> how changes in surrounding l<strong>and</strong>scape will<br />

influence reservoir ecological processes <strong>and</strong> consequently water quality.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes 2010. L<strong>and</strong>scape runoff, precipitation variation <strong>and</strong> reservoir limnology<br />

493<br />

References<br />

Armengol J., Garcia J. C., Comerma M., Romero M., Dolz J., Roura M., Han B. H., Vidal A<br />

<strong>and</strong> Šimek K.,1999. Longitudinal processes in Canyon type reservoirs : The case of Sau<br />

(N. E. Spain). J. G Tundisi & M. Straškraba (Eds.) Theoretical Reservoir Ecology<br />

International Institute of Ecology, Brazilian Academy of Sciences. São Carlos: 313-345.<br />

Geraldes, A.M. <strong>and</strong> Boavida, M. J., 2007. Zooplankton assemblages in two reservoirs: One<br />

subjected to accentuated water levels fluctuations, the other with more stable water<br />

levels. Aquatic Ecology 41:273-284.<br />

Geraldes, A.M. <strong>and</strong> Boavida, M. J., 2004. Limnological variations of a reservoir during two<br />

successive years: One wet, another dry. Lakes <strong>and</strong> Reservoirs: Research & Management.<br />

9:143-152.<br />

Instituto de Meteorologia, 2009. Boletins Climáticos URL: http://www.meteo.pt tecnico-<br />

/pt/publicacoes/cientif/noIM/boletins/index.jspcmbDep=cli&cmbTema=pcl&idDep=cl<br />

i&idTema=pcl&curAno=-1<br />

Soria J. M., Vicente E. <strong>and</strong> Miracle M. R., 2000. The influence of flash floods on the<br />

limnology of the Albufera of Valencia Lagoon (Spain). Verh. internat. Verein.<br />

Limnol. 27: 2232-2235.<br />

500<br />

450<br />

p<br />

re<br />

cip<br />

ita<br />

tio<br />

n<br />

(m<br />

m<br />

)<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

Oct. Nov. Dec. Jan. Feb. Mar. Apr. May Jun. Jul. Aug. Sep.<br />

2000/2001 2001/2002 2007/2008 2008/2009 mean value (1970‐2000)<br />

Figure1: Precipitation noticed for hydrological years of 2000/2001, 2001/2002, 2007/2008 <strong>and</strong> 2008/2009<br />

in the Bragança city (Agroclima Lab-<strong>ESA</strong>B unpubl. data) the other data (Instituto de Meteorologia, 2009)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes 2010. L<strong>and</strong>scape runoff, precipitation variation <strong>and</strong> reservoir limnology<br />

494<br />

2<br />

y<br />

1,5<br />

W0102<br />

1<br />

S0001<br />

Axis 2<br />

0,5<br />

0<br />

S0809 S0102<br />

S0708<br />

W0708<br />

-0,5<br />

W0809<br />

W0001<br />

-1<br />

-1,5<br />

-1<br />

0<br />

Axis 1<br />

Figure 2: N-MDS analysis for crustacean zooplankton data.<br />

1<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes 2010. L<strong>and</strong>scape runoff, precipitation variation <strong>and</strong> reservoir limnology<br />

495<br />

Table 1: Total precipitation, mean <strong>and</strong> st<strong>and</strong>ard deviation (in brackets) of environmental variables<br />

Variables 2000/2001 2001/2002 2007/2008 2008/2009<br />

Winter Summer Winter Summer Winter Summer Winter Summer<br />

Total precipitation (mm) 1482 156.9 423.5 217.7 245.5 263.9 315.9 97.7<br />

TP (μg/l) 66.96 (21.70) 64.30(18.15) 60.29(9.85) 61.45(7.15) 47.53(10.06) 31.56(7.16) 46.67(18.62) 31.67(24.01)<br />

Water temperature (ºC) 10.68 (2.66) 20.33(3.19) 10.09 (3.77) 18.87(4.72) 10.54(3.27) 18.38(3.55) 10.58(3.65) 20.00(4.04)<br />

Conductivity (μS cm -1 ) 54.67(10.67) 58.92(6.65) 48.42(4.92) 64.33(9.17) 91.60 (2.78) 90.35(3.19) 89.92(4.56) 98.78(1.16)<br />

DO (mg/l) 9.06 (1.39) 9.70(1.42) 8.71(0.46) 8.48(0.76) 8.58(1.16) 8.58(1.29) 9.67(0.67) 9.44(1.07)<br />

pH 6.6-7.0 7.0-8.4 6.9-8.3 7.5-7.9 6.0-6.4 6.3-7.2 6.0-6.8 6.5-8.2<br />

Water transparency (m) 2.83(1.51) 4.33(0.44) 3.00(0.71) 4.21(1.74) 4.30(0.67) 5.75(0.42) 5.42(1.07) 6.42(1.5)<br />

Chl a (μg/l) 2.05(0.95) 1.16(0.77) 3.31 (1.68) 1.10(0.48) 2.54(1.46) 1.58(1.33) 1.31(3.45) 2.67(1.50)<br />

Anabaena (ind l -1 x 10 3 ) 0.23(0.26) 14.97(33.76) 90.07(97.91) 0.29(0.43) 0.0 0.0 0.0 4.78(7.84)<br />

Table 2: Mean densities <strong>and</strong> st<strong>and</strong>ard deviation (in brackets) of the crustacean zooplankton species.<br />

Species (ind m 3 x 10 3 ) 2000/2001 2001/2002 2007/2008 2008/2009<br />

Winter Summer Winter Summer Winter Summer Winter Summer<br />

Cladocera<br />

Daphnia longispina/pulex 0.90 (0.98) 1.89(3.72) 0.30(0.19) 0.54(0.69) 0.62(0.71) 2.16(2.07) 3.41(2.51) 1.30(1.69)<br />

Ceriodaphnia pulchella 2.98 (5.79) 3.19(3.70) 0.14 (0.24) 3.93(4.33) 1.19(1.35) 3.52(6.47) 2.32(4.30) 3.60(7.87)<br />

Diaphanosoma. brachyurum 0.0 0.31(0.54) 0.0 0.71(1.13) 0.21(0.31) 1.24(2.11) 0.40(0.83) 1.21(1.16)<br />

Bosmina longirostris 0.07 (0.09) 0.22(0.18) 0.06(0.02) 0.17(0.16) 0.06(0.08) 0.06(0.05) 0.07(0.04) 0.19(0.32)<br />

Copepoda<br />

Acanthocyclops robustus 0.04(0.04) 0.21(0.16) 0.10 (0.12) 0.09(0.66) 0.11(0.06) 0.07(0.04) 0.05(0.07) 0.13(0.14)<br />

Copidodiaptomus numidicus 2.65(2.20) 3.47(2.84) 0.86(0.86) 5.67(0.92) 3.07(2.02) 4.39(1.36) 3.66(3.20) 4.52(1.50)<br />

Nauplii 0.22(0.28) 1.97(1.53) 0.88(0.67) 2.56(1.50) 1.65(1.63) 1.63(0.82) 2.18(2.34) 3.18(3.37)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes & M.J. Boavida 2010. Reservoirs: Mirrors of the surrounding l<strong>and</strong>scape<br />

496<br />

Reservoirs: Mirrors of the surrounding l<strong>and</strong>scape<br />

A.M. Geraldes 1* & M. J. Boavida<br />

1<br />

CIMO, <strong>Escola</strong> <strong>Superior</strong> Agrária do Instituto Politécnico de Bragança Campus de<br />

Santa Apolónia, Bragança, Portugal<br />

2<br />

Centro de Biologia Ambiental, Departamento de Biologia Animal, Faculdade de<br />

Ciências da Universidade de Lisboa Campo Gr<strong>and</strong>e, Portugal<br />

Abstract<br />

To assess in what extent the environmental quality of aquatic systems reflect l<strong>and</strong>scape features<br />

several water quality parameters were determined in two reservoirs. Concomitantly, the<br />

surrounding l<strong>and</strong>scape was characterized <strong>and</strong> the existing potential sources of phosphorous <strong>and</strong><br />

nitrogen runoff were identified <strong>and</strong> when possible estimated. Located in a mountainous area<br />

with negligible direct human influence, it was expected to find lower amounts of suspended<br />

organic matter <strong>and</strong> nutrients in Serra Serrada Reservoir. Water level fluctuations caused by<br />

intensive human water use, grazing <strong>and</strong> frequent l<strong>and</strong> fires in the surrounding l<strong>and</strong>scape can<br />

explain the unexpected high values of the mentioned parameters. In Azibo Reservoir the factors<br />

with greatest influence on water parameters seem to be allochthonous sources of nutrients<br />

originated from agriculture <strong>and</strong> grazing in the catchment area <strong>and</strong> recreational activities.<br />

However, in this particular case the potential sources of nutrients could have been minimized by<br />

the patchy structure of the surrounding l<strong>and</strong>scape.<br />

Keywords: L<strong>and</strong>scape, Water quality L<strong>and</strong> use, Phosphorous <strong>and</strong> Nitrogen sources<br />

1. Introduction<br />

Pressure caused by human activities in the catchment area <strong>and</strong> reservoir vicinity generally leads<br />

to an intensification of surface runoff, causing an increase in eutrophication, thus threatening<br />

water quality. Runoff rates depend mainly on l<strong>and</strong> use, vegetation cover <strong>and</strong> l<strong>and</strong>scape mosaic.<br />

The present study was carried out for two years on S. Serrada <strong>and</strong> Azibo reservoirs located at<br />

the Portuguese part of the international Douro catchment basin, in Trás-os-Montes region (NE<br />

Portugal). Location, morphological <strong>and</strong> hydrological characteristics of both reservoirs are<br />

shown in Table 1. Serra Serrada Reservoir was built to supply water to the city of Bragança <strong>and</strong><br />

to generate hydroelectric power. Consequently, pronounced water level fluctuations occur,<br />

ranging between 8 <strong>and</strong> 10 m. Direct human influence on this reservoir impoundment is<br />

considered negligible. There are no villages, there has been no agricultural activity for<br />

approximately 20 years <strong>and</strong> recreational activities are not significant. However, in the catchment<br />

basin grazing can be very intense in summer months. All over the year there are only about 200<br />

sheep grazing in the S. Serrada catchment, yet from May to August about 5,000 sheep from<br />

lowl<strong>and</strong>s are transported from the surrounding lowl<strong>and</strong>s to graze in the catchment <strong>and</strong> reservoir<br />

surroundings. Consequently, this area is very often subjected to wildfires which are mainly<br />

induced by shepherds to obtain better graze. Azibo Reservoir was built for water supply <strong>and</strong><br />

irrigation, but those activities are not significant <strong>and</strong> the reservoir is used mainly for recreation.<br />

In Azibo direct influence of human activities is greater during summer when reservoir <strong>and</strong><br />

* Corresponding author. Tel.:351273303217 - Fax:351273331570<br />

Email address: geraldes@ipb.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes & M.J. Boavida 2010. Reservoirs: Mirrors of the surrounding l<strong>and</strong>scape<br />

497<br />

surroundings are used by about 10,000 people for recreation such as swimming, camping <strong>and</strong><br />

boating. Angling is also an important activity. The watershed area is occupied by meadows<br />

(1286 ha), woodl<strong>and</strong>s <strong>and</strong> scrub (935 ha) <strong>and</strong> extensive agriculture (2235 ha). The latter is<br />

found all over the year <strong>and</strong> the main crops are olives (657 ha), chestnuts (650 ha), cereals (546<br />

ha), vineyards (144 ha) <strong>and</strong> potato (85 ha). In the reservoir shore there are intensive crops,<br />

which are less than 1% of all crops (INE, 1999). Extensive grazing also occurs in this drainage<br />

basin. In Azibo catchment area there are several small villages. The total of inhabitants is about<br />

1,500 <strong>and</strong> most of them are more than 50 years old (INE, 2001). Passing nearby <strong>and</strong> over of one<br />

stream that feed this reservoir there is a highway, IP4. In this highway the average daily traffic<br />

volume is 6,000 vehicles (Barbosa <strong>and</strong> Hvitved-Jacobsen, 1999). There is no industrial activity<br />

in both reservoir catchments. Therefore, the purpose of this research was to find out whether in<br />

what extent the environmental quality of aquatic systems reflect l<strong>and</strong>scape occupation by<br />

assessing several water quality parameters.<br />

2. Methodology<br />

Water samples were collected monthly in winter <strong>and</strong> biweekly in summer, from January 2000 to<br />

December 2001 in both reservoirs. Details concerning sampling methodology, laboratorial<br />

procedure, statistical analysis <strong>and</strong> the determination of potential allochthonous sources of<br />

phosphorus <strong>and</strong> nitrogen can be found in Geraldes <strong>and</strong> Boavida (2003).<br />

3. Results<br />

Maximum <strong>and</strong> minimum ranges as well as mean <strong>and</strong> st<strong>and</strong>ard deviation values for water<br />

temperature, dissolved oxygen, conductivity, pH <strong>and</strong> water colour observed for both S. Serrada<br />

<strong>and</strong> Azibo are presented on Table 2. Both reservoirs were classified as meso-eutrophic. The<br />

potential allochthonous sources of phosphorus <strong>and</strong> nitrogen are presented in figure 1. In S.<br />

Serrada, grazing can contribute 30,450 kg of N <strong>and</strong> 13,050 kg of P per year. Wildfires can also<br />

contribute with substantial loads. However, for this region there are no data allowing<br />

quantification of this source. In Azibo trophic state is possibly influenced by agriculture,<br />

grazing, sewage, as well as by angling <strong>and</strong> bathing. Agriculture can contribute 36,394 kg of N<br />

<strong>and</strong> 49,192 kg of P per year, grazing 172,956 kg of N <strong>and</strong> 60,786 kg of P per year <strong>and</strong> sewage<br />

4,927 kg of N <strong>and</strong> 1,752 Kg of P per year. Angling <strong>and</strong> bathing can contribute 28,080 kg of N<br />

<strong>and</strong> 5,184 kg of P <strong>and</strong> 900 kg of N <strong>and</strong> 41.5 kg of P per year, respectively. Results of MDS are<br />

depicted on figure 2. In this kind of diagram sample to sample dissimilarity is represented by the<br />

distance between points. So, the gradation in the spread of the points indicated the existence of<br />

two groups: One formed by samples obtained at S. Serrada <strong>and</strong> one formed those obtained at<br />

Azibo. This means that there is an inter-reservoirs variability related with some of the studied<br />

parameters. In fact, according to Kolmogorov - Smirnov test, nutrient <strong>and</strong> CHL a concentrations<br />

showed no significant differences between reservoirs in spite of differences in l<strong>and</strong>scape<br />

occupation, water use <strong>and</strong> exposure to different degrees of disturbance. However, differences<br />

between reservoirs were found for conductivity (D m = 1; P


A.M. Geraldes & M.J. Boavida 2010. Reservoirs: Mirrors of the surrounding l<strong>and</strong>scape<br />

498<br />

disturbance. The observed differences in water temperature, conductivity <strong>and</strong> pH might be the<br />

result of the synergistic effect of reservoir altitude, <strong>and</strong> geological zone.<br />

S. Serrrada is a highly disturbed system if compared with other located in similar geological <strong>and</strong><br />

climate regions (Negro et al., 2001). There are two sorts of disturbance: One internal because of<br />

water level fluctuation, <strong>and</strong> the other external originating by the combined effect of grazing <strong>and</strong><br />

fire. Furthermore, the areas of both reservoir <strong>and</strong> catchment are small. Thus, the intensity of use<br />

of water <strong>and</strong> grazing activity which could not have had a strong impact in systems with larger<br />

areas, in this reservoir could have induced a severe reduction in water quality. In fact, the<br />

observed ressuspension of bottom sediments <strong>and</strong> organic matter following water runoff input<br />

after the first rains, plus the turbulence generated by water level rising <strong>and</strong> the periodical<br />

exposure of littoral sediments to cycles of drying <strong>and</strong> wetting could explain the high phosphorus<br />

concentrations. On the other h<strong>and</strong>, grazing is not only a source of nutrients but also the main<br />

cause of fires in this region, since those are induced by shepherds to obtain better graze.<br />

Actually, this catchment is one of the areas with more fires per year in the Montesinho Natural<br />

Park (Rainha <strong>and</strong> Cabral, 2001). There are no studies quantifying the nutrient inputs to this<br />

reservoir because of the erosive impact of rainfall on post-burned soils. However, considering<br />

the high slope <strong>and</strong> the dominant soil type in this area, which according to Agroconsultores <strong>and</strong><br />

Coba (1991) bear a high potential risk of erosion, high rates of soil erosion <strong>and</strong> consequently<br />

high surface runoff are expected. Besides, some research developed in other regions has shown<br />

that the consequences of a fire can be the increase in trophic state <strong>and</strong> the subsequent decrease<br />

of water quality in the adjacent water bodies. Those effects are more accentuated in sloppy areas<br />

<strong>and</strong> after intense precipitation events.<br />

In Azibo internal disturbance caused by water level fluctuation is minimal. Moreover,<br />

the areas of reservoir <strong>and</strong> catchment are larger, l<strong>and</strong>scape is patchy <strong>and</strong> fires are not frequent.<br />

However, other sources of disturbance such as agriculture, mainly the intensive cultures in<br />

reservoir shore, grazing <strong>and</strong> recreational activities, can explain the trophic state classification.<br />

According to estimations of the potential allochthonous sources of nutrients, agriculture <strong>and</strong><br />

grazing seemed to be the greatest sources of N <strong>and</strong> P in the Azibo catchment. The intensity of<br />

exportation of nutrients from those activities <strong>and</strong> from sewage seems to be highly seasonal. In<br />

fact, in the beginning of the wet season the nutrient concentrations in water runoff were higher<br />

than in the water runoff generated by end of season rains. However, agricultural <strong>and</strong> grazing<br />

sources of nutrients can decrease within few years, since most of the farmers are more than 50<br />

years old nowadays (INE, 1999; 2001) <strong>and</strong> that there is a considerable tendency for human<br />

desertification because of the low rentability of agricultural practices. Besides, the l<strong>and</strong>scape in<br />

this catchment is very patchy <strong>and</strong> consequently there are numerous buffer areas such as<br />

woodl<strong>and</strong>s, meadows <strong>and</strong> riparian vegetation that can minimise those potential sources of<br />

nutrients. Thus, intensive agriculture practices in the reservoir shore <strong>and</strong> recreational activities<br />

in summer are or might become in a near future the main nutrient sources to Azibo Reservoir. In<br />

fact, the values of nutrient concentrations obtained in summer can be related to those activities,<br />

which are more intensive in this period. Another, a possible source of pollutants that can also<br />

affect the water quality in Azibo is the highway IP4. According to Barbosa <strong>and</strong> Hvitved-<br />

Jacobsen (1999) the average concentration levels of Pb, Zn <strong>and</strong> Cu in the IP4 highway runoff<br />

are 10.8, 172 <strong>and</strong> 10.7 µg/l, respectively. Besides, the projected golf course in Azibo Reservoir<br />

shores if implemented might be a important new source of phosphorous <strong>and</strong> nitrogen. From the<br />

obtained data it seems undeniable that reservoirs can be regarded as mirrors of the surrounding<br />

l<strong>and</strong>scape. However, there is a lack of data concerning for instance soil nutrient retention<br />

capacity <strong>and</strong> erosion rates. Such data are fundamental to develop export coefficient models<br />

adapted to these areas, allowing the correct estimation of nutrient <strong>and</strong> pollutant inputs, <strong>and</strong> to<br />

make possible the development of correct management measures for these reservoirs <strong>and</strong> the<br />

surrounding l<strong>and</strong>scape.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes & M.J. Boavida 2010. Reservoirs: Mirrors of the surrounding l<strong>and</strong>scape<br />

499<br />

References<br />

Agroconsultores <strong>and</strong> Coba 1991. Carta de Solos. Projecto de Desenvolvimento Rural Integrado<br />

de Trás-os-Montes. UTAD, Vila Real.<br />

Barbosa A. E. <strong>and</strong> Hvitved-Jacobsen,T. 1999. Highway runoff <strong>and</strong> potential for removal of<br />

heavy metals in an infiltration pond in Portugal. Sci. Tot. Environ. 235: 151-159.<br />

Geraldes, A.M. <strong>and</strong> Boavida, M. J., 2003. Distinct age <strong>and</strong> l<strong>and</strong>scape influence on two<br />

reservoirs under the same climate. Hydrobiologia 504:277-288<br />

INE, 1999. Censos da Agricultura 1999. Instituto Nacional de Estatística.<br />

INE, 2001. Censos da População 2001. Instituto Nacional de Estatística.<br />

Negro, A. I., C. De Hoyos, C. <strong>and</strong> Vega, J. C., 2000. Phytoplankton structure <strong>and</strong> dynamics in<br />

Lake Sanabria <strong>and</strong> Valparaíso reservoir (NW Spain). Hydrobiologia 424: 25-37.<br />

Rainha, M. <strong>and</strong> Cabral, P., 2001. Incêndios ocorridos na área do Parque Natural de Montesinho<br />

(1994-2001). Parque Natural de Montesinho (Internal Report)<br />

Table 1: Main general features of S. Serrada <strong>and</strong> Azibo reservoirs.<br />

S.Serrada<br />

Azibo<br />

Location<br />

Latitude: 41º57’12’’(N)<br />

Longitude: 6º 46’ 44’’ (W)<br />

Latitude: 41º32’50’’(N)<br />

Longitude: 6º 53’ 38’’ (W)<br />

Altitude (m) 1300 500<br />

Geology granitic bedrock schistic bedrock<br />

Mean annual precipitation (mm) 1300 800-1000<br />

Watershed area (Km 2 ) 6.7 89.0<br />

Reservoir area (Km 2 ) 0.25 4.10<br />

Total capacity (m 3 ) 1680 x 10 3 54470 x 10 3<br />

Max. Depth (m) 18 30<br />

Mean depth (m) 6.72 13.2<br />

Water residence time (years) 0.36 2.22<br />

Year of filling 1995 1982<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes & M.J. Boavida 2010. Reservoirs: Mirrors of the surrounding l<strong>and</strong>scape<br />

500<br />

Table 2: Physical factors recorded for S. Serrada <strong>and</strong> Azibo reservoirs. Are shown minimum-maximum<br />

range; mean <strong>and</strong> st<strong>and</strong>ard deviation in brackets.<br />

2000 2001<br />

S. Serrada<br />

Water temperature (ºC) 1.46-21.39 (12.51/6.59) 2.70-20.19 (12.90/6.46)<br />

Dissolved oxygen (mg/l) 7.38-12.42 (8.55/1.45) 6.20-10.72 (8.55/1.49)<br />

Conductivity (µS/cm) 4-10 (6.94/1.84) 3-8 (5.95/1.61)<br />

Water transparency (m) 4.50-1.00 (2.66/1.09) 1-5 (2.85/1.14)<br />

pH 5.77-6.56 (6.18/0.28) 5.95-8.34 (6.89/1.03)<br />

Water colour (Pt units) 0.44-46.35 (17.39/14.27) 5.24-35.45 (18.55/9.36)<br />

Azibo<br />

Temperature (ºC) 5.58-24.7 (16.52/6.13) 8.06-23.83 (17.16/5.73)<br />

Dissolved oxygen (mg/l) 7.54-11.53 (9.09/1.06) 7.21-10.78 (8.99/1.33)<br />

Conductivity (µS/cm) 51-81 (69.87/11.09) 43-66 (56.23/7.64)<br />

Water transparency (m) 1.5-6.0 (4.72/1.25) 1.5-5.5 (3.50/1.14)<br />

pH 6.71-8.05 (7.33/0.37) 6.64-8.36 (7.40/0.81)<br />

Water colour (Pt units) 0.0-9.02 (2.82/3.02) 0.0-21.81 (6.21/5.66)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.M. Geraldes & M.J. Boavida 2010. Reservoirs: Mirrors of the surrounding l<strong>and</strong>scape<br />

501<br />

kg/year x 10 3<br />

200<br />

160<br />

120<br />

80<br />

40<br />

B<br />

0<br />

Agriculture<br />

Grazing<br />

Sewage<br />

Angling<br />

Bathing<br />

Wildfires<br />

N<br />

P<br />

Figure 1: Potential external sources of nitrogen <strong>and</strong> phosphorus to S. Serrada (A) <strong>and</strong> Azibo (B)<br />

reservoirs ( means non quantified source).<br />

Figure 2: Results of MDS ordination (ss-Samples performed in S. Serrada; az-Samples performed in<br />

Azibo).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. Koschke et al. 2010. Using a multi-criteria approach to fit the evaluation basis of “Pimp Your L<strong>and</strong>scape”<br />

502<br />

Using a multi-criteria approach to fit the evaluation basis of the<br />

modified 2-D cellular automaton “Pimp Your L<strong>and</strong>scape”<br />

Lars Koschke * , Christine Fürst, Carsten Lorz, Susanne Frank & Franz Makeschin<br />

Institute for Soil Science <strong>and</strong> Site Ecology, Dresden University of Technology,<br />

Germany<br />

Abstract<br />

This contribution presents an evaluation approach that is used within the framework of the l<strong>and</strong>use<br />

management support tool “Pimp your l<strong>and</strong>scape”. “Pimp your l<strong>and</strong>scape” is a modified 2-D<br />

cellular automaton, which enables a multi-criteria impact assessment of l<strong>and</strong>-use management<br />

planning decisions. To evaluate l<strong>and</strong>-cover changes, values are assigned to each l<strong>and</strong>-use type;<br />

these values describe its contribution to regionally specific sets of ecosystem services <strong>and</strong> / or<br />

environmental risks.<br />

A major problem is that universally valid indicators that facilitate the estimation of such values<br />

are rare. Indicators taken from sectoral evaluation approaches, such as those that are available<br />

from forest or agricultural research, often times address different target scales, thus negating<br />

their suitability for a holistic evaluation on l<strong>and</strong>scape level.<br />

We focus here on the question of how to quantify the impact of planning measures. A<br />

combination of participatory multi-criteria <strong>and</strong> indicator-based analysis is developed to arrive at<br />

a comprehensive evaluation.<br />

Keywords: Pimp Your L<strong>and</strong>scape, ecosystem services, l<strong>and</strong>-use type, l<strong>and</strong>scape evaluation,<br />

criteria <strong>and</strong> indicators<br />

1. Introduction<br />

L<strong>and</strong>-use pattern change is one of many important factors that generate climate change. L<strong>and</strong>management<br />

affects ecosystem functioning in many ways. Therefore, management measures<br />

<strong>and</strong> planning targets connected to climate change adaption are related to alterations in l<strong>and</strong>-use<br />

types or management intensity. These changes will impact ecosystem services (MEA 2005). For<br />

the current project region in Saxony, Germany, we have chosen six ecosystem services:<br />

ecological functioning, aesthetical value, human health <strong>and</strong> well being, CC mitigation, bioresource<br />

provision <strong>and</strong> economic wealth.<br />

In our approach we aim at estimating the consequences of l<strong>and</strong> cover <strong>and</strong> management changes<br />

on l<strong>and</strong>scape level. We intend to support decision makers who are confronted with the challenge<br />

to take into account multiple objectives. For this purpose the tool Pimp your l<strong>and</strong>scape (PYL)<br />

has been developed. PYL is software that simulates <strong>and</strong> visualizes the impact of management or<br />

planning scenarios on ecosystem services at the l<strong>and</strong>scape level; these include structural<br />

l<strong>and</strong>scape aspects <strong>and</strong> environmental data (climate, soil, topography). The end-user receives<br />

feed-back in real-time <strong>and</strong> can easily simulate <strong>and</strong> compare manifold scenarios. The software is<br />

based on a cellular automaton approach including GIS-features. The cell size is fixed at 100 x<br />

100m. To each cell, the predominant l<strong>and</strong>-use type (LUT) taken from the CORINE l<strong>and</strong>-cover<br />

(CLC) 2000 or regional l<strong>and</strong> cover data is assigned (Fürst et al. 2010).<br />

* Corresponding author. Tel.: +49 35203 38 31377 - Fax: +49 35203 38 31388<br />

Email address: Lars.Koschke@tu-dresden.de<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. Koschke et al. 2010. Using a multi-criteria approach to fit the evaluation basis of “Pimp Your L<strong>and</strong>scape”<br />

503<br />

2. Methodology<br />

Within PYL the provision of ecosystem services is evaluated in a first step for each l<strong>and</strong>-use<br />

type <strong>and</strong> infrastructural element. In a second step, the value at the l<strong>and</strong>scape level is calculated<br />

as the average of all cell values; this includes an additional correction of the result from the<br />

point of view of positive or negative aspects gleaned from the l<strong>and</strong>scape structure (cf. Fig. 1 in<br />

Fürst et al., 2010).<br />

Within PYL the l<strong>and</strong> use classification st<strong>and</strong>ards of CLC 2000 <strong>and</strong> the environmental services<br />

(<strong>and</strong> functions) (LUF) set previously described (Pérez-Soba et al. 2008) are available as initial<br />

settings. The user can modify these initial settings or adopt completely different settings<br />

according to the regional application targets. After having selected a set of LUTs <strong>and</strong> LUFs, the<br />

resulting value matrix must be filled out. Evaluation of the l<strong>and</strong>-use types follows classification,<br />

on a relative scale from 0 (worst case) to 100 (best case), of the specific regional contribution of<br />

a LUT to a LUF. The introduction of this relative scale aims to facilitate the multi-criteria<br />

evaluation of planning measures.<br />

The value matrix contains initial impact values of the l<strong>and</strong> use types <strong>and</strong> infrastructural elements<br />

on the environmental services. The initial value of a l<strong>and</strong> use type for an environmental service<br />

represents the maximum in the regional context, which can only be reduced (a) with regard to<br />

environmental cell attributes from additional information layers such as height above sea level,<br />

mean annual precipitation, temperature, soil type <strong>and</strong> exposition. (b) The impact of the cell<br />

environment (homogeneous l<strong>and</strong> use types vs. different l<strong>and</strong> use types) <strong>and</strong> neighbourhood type<br />

(edge to edge vs. corner to corner) can impact the original maximum value. The influence of<br />

structural features on the final evaluation is discussed elsewhere (cf. contribution of<br />

Frank et al.).<br />

2.1 Indicator-based assessment<br />

2.1.1 Statistical indicators<br />

Evaluation of ecosystem services is commonly based on indicators. For our assessment the<br />

challenge is to find or generate indicators associated to LUTs which can then be processed in an<br />

aggregation framework to estimate comparative scores.<br />

Despite the general scarcity of indicators, a few economic-evaluation indicators are available<br />

that allow for a comparison of LUT-related l<strong>and</strong> prices. We used st<strong>and</strong>ard ground value, l<strong>and</strong><br />

rent <strong>and</strong> market price to compute l<strong>and</strong>-use dependent value points. We had two prerequisites for<br />

the choice of an indicator. The first condition was a thematic relation to at least two LUTs.<br />

Second, data had to be available on a per unit area basis due to 100 x 100m grid cells used as<br />

the reference unit in PYL. Quotients of indicator values were calculated to reflect the value of<br />

an LUT in comparison to another LUT (see table 1). For each LUT the mean of quotient-values<br />

was determined <strong>and</strong> normalised to the scoring scale (0 – 100 value points) using equation 1. The<br />

st<strong>and</strong>ard ground value acts as a “cross link” as it integrates information about semi-natural <strong>and</strong><br />

artificial (urban) LUTs.<br />

Tab. 1<br />

2.1.2 Measured <strong>and</strong> modeled indicators<br />

Other sources of indicators are regional studies with a l<strong>and</strong>-use oriented background. Data<br />

relating to N-export (LfULG 2009), soil sealing <strong>and</strong> run-off coefficient (Arlt 2001) could be<br />

derived from such studies (table 2). Indicator values were normalized using equation 1 or 2.<br />

When an indicator’s link to the performance of an LUF was negative, equation 2 was utilized. In<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. Koschke et al. 2010. Using a multi-criteria approach to fit the evaluation basis of “Pimp Your L<strong>and</strong>scape”<br />

504<br />

this concept, an indicator can be applied to one or more LUFs. Since the impact of an indicator<br />

on an LUF may vary, weights need to be indicated.<br />

(1)<br />

(2)<br />

where is the indicator value for a given LUT normalised to a score between 0 <strong>and</strong> 100<br />

<strong>and</strong> is the value of the indicator assigned to the LUT. <strong>and</strong> correspond to the<br />

minimum <strong>and</strong> maximum of indicator values.<br />

Tab. 2<br />

2.2 Participatory-based assessment<br />

A preliminary study showed that l<strong>and</strong>scape assessment based on an evaluation of LUTs cannot<br />

be achieved by using quantitative indicators only (cf. chapter 2.1). With measured or calculated<br />

indicator values most evaluation-criteria cannot be assessed across l<strong>and</strong>-uses. For this, CLC<br />

(2000) classes are too coarse in terms of spatial <strong>and</strong> thematic resolution. At the l<strong>and</strong>scape level a<br />

more general approach is required.<br />

Multi-Criteria Analysis (MCA) involving expert consultation is a suitable tool for support of the<br />

evaluation. Because of their comprehensive character, MCA techniques are often used to rank<br />

alternatives for decision-making or to assign values, if there are multiple objectives involved.<br />

LUFs were broken down to a set of criteria which are actually evaluated. Table 3 illustrates a<br />

draft of a matrix showing LUFs (environmental services) <strong>and</strong> connected criteria which are<br />

relevant in the context of the REGKLAM-project. LUTs are placed on the y-axis, while<br />

environmental services are arranged on the x-axis. The matrix will be used to appraise the<br />

capacity of the LUTs to provide goods <strong>and</strong> services.<br />

Tab. 3<br />

We combine normalized indicator values <strong>and</strong> expert estimation within the matrix. The expert<br />

group should (i) discuss relevant functions <strong>and</strong> criteria, <strong>and</strong> (ii) interpolate gaps between<br />

indicator values for LUTs that have not been ascertained. In addition (iii) the respondents will<br />

be asked to evaluate the LUTs in terms of other criteria, the description of which cannot be<br />

related to any measurable indicator values.<br />

The contribution of an indicator depends on regional significance <strong>and</strong> the number of indicators/<br />

criteria that correspond to an LUF. Hence, prior to an aggregation framework, stakeholders (iv)<br />

have to assign weights to the criteria, whereby preferences are expressed as well. According to<br />

the preliminary matrix, experts would have to make 529 decisions.<br />

3. Results <strong>and</strong> Discussion<br />

Indicator-based assessment of capacities of LUTs to provide certain services has proven<br />

difficult. When using an indicator set for evaluation of the economic value, an extreme<br />

imbalance between the value of urban <strong>and</strong> semi-natural (rural) areas results. As a consequence,<br />

for example, the contribution of forest <strong>and</strong> agricultural areas to the regional economy is mostly<br />

underestimated. In addition, the utilisation of indicators which are highly aggregated is critical.<br />

L<strong>and</strong> rent <strong>and</strong> market value are indirectly contained in the st<strong>and</strong>ard ground value, in the<br />

calculation of which market-driven factors are also considered. The scores of table 1 reflect the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. Koschke et al. 2010. Using a multi-criteria approach to fit the evaluation basis of “Pimp Your L<strong>and</strong>scape”<br />

505<br />

disproportionality of surface share <strong>and</strong> marketed economic value of LUTs dominated by<br />

agricultural or silvicultural use in comparison to built-up areas. This discrepancy is confirmed<br />

by sectoral accountings.<br />

The use of modelled <strong>and</strong> measured indicators is very restricted, not only regarding economic<br />

aspects. Data about N-Export, soil sealing <strong>and</strong> run-off coefficient as estimated for different<br />

LUTs in other studies can be used in the evaluation matrix (table 3). N-export <strong>and</strong> soil sealing<br />

act as proxies for l<strong>and</strong>-use intensity <strong>and</strong> the functioning of ecological processes. Soil sealing<br />

also has an impact on l<strong>and</strong>scape aesthetics, assuming that a high share of sealed surface area<br />

causes a decreasing visual attractiveness. Run-off coefficient is connected to water retention<br />

potential which is important for the risk of flooding during heavy rainfalls. Therefore this<br />

indicator is considered as a criterion for CC mitigation. For assessing ecological functioning<br />

species richness, percentage of dead wood or other non-structural indicators are not convenient.<br />

Depending on environmental conditions <strong>and</strong> habitat composition, such indicators behave<br />

differently <strong>and</strong> are thus not appropriate to be assigned to LUTs.<br />

The drawbacks <strong>and</strong> restrictions related to the integration of indicators of multiple scales <strong>and</strong><br />

dimensions resulted in the consideration of stakeholder involvement to obtain a comprehensive<br />

evaluation. Participatory methods are often considered the most suitable approach in terms of<br />

l<strong>and</strong>scape analysis (eg. de Groot 2006). They have been widely <strong>and</strong> successfully applied in<br />

l<strong>and</strong>scape/ environmental modelling (Dai et al. 2001; Vacik <strong>and</strong> Lexer 2001). Due to weighting<br />

<strong>and</strong> aggregation issues, indicator-based approaches <strong>and</strong> collaborating techniques are both prone<br />

to biases. MCA approaches also suffer from concerns such as the selection of participants, their<br />

respective background <strong>and</strong> interests (Danae <strong>and</strong> Stelios 2004). We therefore intend to<br />

incorporate available indicator values as reference points for experts during the evaluation<br />

process.<br />

4. Conclusion<br />

Some pre-analysis in our work proved that a purely indicator based approach for a holistic<br />

l<strong>and</strong>scape evaluation in Pimp your l<strong>and</strong>scape will fail, because fitting indicators are rare <strong>and</strong><br />

sectoral indicator approaches are not suitable for up-scaling the evaluation results (Fürst et al.,<br />

subm.). A comprehensive evaluation of LUTs in the context of l<strong>and</strong>scape management <strong>and</strong><br />

planning was founded on several data sources. We suggest a mixed indicator-based <strong>and</strong> expert<br />

opinion evaluation approach.<br />

Next steps are expert-evaluation which may be followed up by another round in order to<br />

estimate the performance of LUTs under future scenarios, which is important for the<br />

introduction of time slots. Thus, based on CC projections also ecosystem dynamics can be<br />

regarded. Since we de facto assess l<strong>and</strong>-cover types as supported by CLC (2000) which are<br />

related to but not fully congruent with LUTs, an incorporation of l<strong>and</strong> management types into<br />

PYL is planned in order to address the impact of adaptation options to CC at the farm level<br />

(irrigation, crop rotation, tillage, changing tree species composition etc.).<br />

Despite several shortcomings <strong>and</strong> limitations inherent in the approach, we believe that<br />

comprehensive assessments are of great importance for environmental managers <strong>and</strong> for the<br />

consideration of ecosystem services in l<strong>and</strong>scape planning processes.<br />

Reference<br />

Arlt, G. 2001. Auswirkungen städtischer Nutzungsstrukturen auf Bodenversiegelung und<br />

Bodenpreis, pp. 183. Institut für Ökologische Raumentwicklung, Dresden.<br />

Dai, F.C., Lee, C.F., <strong>and</strong> Zhang, X.H. 2001. GIS-based geo-environmental evaluation for urban<br />

l<strong>and</strong>-use planning: a case study. Engineering Geology 61: 257-271.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. Koschke et al. 2010. Using a multi-criteria approach to fit the evaluation basis of “Pimp Your L<strong>and</strong>scape”<br />

506<br />

Danae, D., <strong>and</strong> Stelios, G. 2004. Multicriteria analysis. European Commission Externalities of<br />

Energy Research Network Working Paper. Greece: National Technical University<br />

Athens.<br />

de Groot, R. 2006. Function-analysis <strong>and</strong> valuation as a tool to assess l<strong>and</strong> use conflicts in<br />

planning for sustainable, multi-functional l<strong>and</strong>scapes. L<strong>and</strong>scape <strong>and</strong> Urban Planning<br />

75: 175-186.<br />

Fürst, C., Lorz, C., Pietzsch, K., Koschke, L., Frank, S., <strong>and</strong> Makeschin, F. 2010. Pimp your<br />

l<strong>and</strong>scape – a cellular automaton approach to estimate the effects of l<strong>and</strong> use pattern<br />

changes on environmental services. Proceedings of the International Conference on<br />

Integrative L<strong>and</strong>scape Modelling LANDMOD 2010. .<br />

LfULG. 2009. Atlas der Nährstoffeinträge in sächsische Gewässer. Sächsisches L<strong>and</strong>esamt für<br />

Umwelt L<strong>and</strong>wirtschaft und Geologie (LfULG).<br />

MEA. 2005. Ecosystems <strong>and</strong> human well-being: Synthesis. A Report of the Millenium<br />

Ecosystem Assessement. . Isl<strong>and</strong> Press, Washington, pp. 155.<br />

Pérez-Soba, M., Petit, S., Jones, L., Bertr<strong>and</strong>, N., Briquel, V., Omodei-Zorini, L., Contini, C.,<br />

Helming, K., Farrington, J.H., Mossello, M.T., et al. 2008. L<strong>and</strong> use functions - a<br />

multifunctionality approach to assess the impact of l<strong>and</strong> use changes on l<strong>and</strong> use<br />

sustainability. In Sustainability Impact Assessment of L<strong>and</strong> Use <strong>Change</strong>s. (eds. K.<br />

Helming, M. Pérez-Soba, <strong>and</strong> P. Tabbush), pp. 375-404. Springer, Berlin-Heidelberg.<br />

Vacik, H., <strong>and</strong> Lexer, M.J. 2001. Application of a spatial decision support system in managing<br />

the protection forests of Vienna for sustained yield of water resources. <strong>Forest</strong> Ecology<br />

<strong>and</strong> Management 143: 65-76.<br />

Tables<br />

Table 1: Simple relational matrix of cross-linked indicators (vertical bars) for assessing the economic<br />

wealth. Quotients of initial indicator values (vertical column on the left) are depicted in the matrix. They<br />

are assigned to a limited number of LUTs. Quotients allow for comparison of l<strong>and</strong> prices among LUTs.<br />

The mean value of each LUT was normalised to a score between 0 (least valuable) <strong>and</strong> 100 (most<br />

valuable) (grey row at the bottom).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


K. Koschke et al. 2010. Using a multi-criteria approach to fit the evaluation basis of “Pimp Your L<strong>and</strong>scape”<br />

507<br />

Table 2: Compilation of data derived from different studies with corresponding value points assigned to<br />

LUTs. LUTs for which no data were available are not depicted.<br />

L<strong>and</strong>-use function (LUF)<br />

Ecological<br />

Functioning<br />

Ecological<br />

functioning/Aesthetics<br />

CC Mitigation<br />

Associated indicator<br />

N-export 1 Soil sealing 2 Run-off<br />

coefficient 2<br />

(kg N ha -1 a -1 ) Points (%) Points (Ψ) Points<br />

Continuous urban fabric 20 0 73 0 0.7 13<br />

Discontinuous urban fabric - ‐ 31 58 0.325 59<br />

Road <strong>and</strong> rail networks - ‐ - - 0.8 0<br />

Mineral extraction sites 9.1 98 - - - -<br />

Non-irrigated arable l<strong>and</strong> - ‐ 0 100 0.15 81<br />

Vineyards 11.2 79 0 100 - -<br />

Fruit trees & berry plantations 11.3 78 0 100 - -<br />

Pastures - ‐ 0 100 0.1 88<br />

Agro-forestry areas - ‐ 0 100 - -<br />

Broad-leaved forest 8.9 100 0 100 - -<br />

Coniferous forest 13.7 57 0 100 - -<br />

Mixed forest - ‐ 0 100 -<br />

Natural grassl<strong>and</strong>s - ‐ 0 100 - -<br />

Water bodies 15.9 37 0 100 0 100<br />

1 (LfULG 2009), 2 (Arlt 2001)<br />

Table 3: Draft matrix for the assessment of regionally relevant l<strong>and</strong>-cover types. Displayed indicator<br />

scores (light grey <strong>and</strong> grey fields) st<strong>and</strong> for an LUTs capacity to sustain specific ecosystem services.<br />

Original statistical <strong>and</strong> measured (modelled) data values were normalized to the scoring scale (0-100). For<br />

incommensurable criteria, values will be estimated <strong>and</strong> existing data gaps (dark grey fields) completed<br />

using expert knowledge. The weighted mean of criteria scores corresponds to the final evaluation <strong>and</strong> will<br />

be displayed in the black fields.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Assessing “spatially explicit” l<strong>and</strong> use/cover change models<br />

508<br />

Assessing “spatially explicit” l<strong>and</strong> use/cover change models<br />

Jean-François Mas 1* , Azucena Pérez Vega 2 & Keith Clarke 3<br />

1<br />

Centro de Investigaciones en Geografía Ambiental - Universidad Nacional Autónoma<br />

de México, Mexico<br />

2<br />

Departamento de Ingeniería Civil - Universidad de Guanajuato, Mexico<br />

3<br />

Department of Geography – University of California - Santa Barbara, USA<br />

Abstract<br />

Spatially explicit l<strong>and</strong> use/cover change (LUCC) models aim at predicting the location <strong>and</strong><br />

pattern of LUCC. The simulation involves a spatial procedure which identifies the potential<br />

locations of change <strong>and</strong> eventually replicates the patterns of the l<strong>and</strong>scape. Generally,<br />

evaluation is based upon the comparison of the simulated map <strong>and</strong> a true map of the same date.<br />

However, most of the evaluation techniques only evaluate the spatial coincidence between<br />

simulated <strong>and</strong> true changes <strong>and</strong> do not assess the ability of the model to simulate the l<strong>and</strong>scape<br />

patterns. Simulated maps obtained by two models (DINAMICA <strong>and</strong> L<strong>and</strong> <strong>Change</strong> Modeler)<br />

were evaluated using a fuzzy similarity index <strong>and</strong> l<strong>and</strong>scape metrics. Results show that more<br />

realistic simulated l<strong>and</strong>scape are often obtained at the expense of location coincidence. When<br />

patterns of l<strong>and</strong>scape is an important issue (e.g. Fragmentation), indices taking into account<br />

spatial patterns, <strong>and</strong> not just location, should be used to assess model performance.<br />

Keywords:l<strong>and</strong> use/cover change, modeling, l<strong>and</strong>scape metrics, assessment<br />

1. Introduction<br />

Over the last decades, a range of “spatially explicit” computational models of LUCC have been<br />

developed for the projection of alternative scenarios into the future, for conducting experiments<br />

that test our underst<strong>and</strong>ing of key processes, <strong>and</strong> for describing the latter in quantitative terms<br />

(Veldkamp <strong>and</strong> Lambin 2001; Xiang <strong>and</strong> Clarke 2003). Among these models, process-based<br />

models, closely related with geographic information systems, view l<strong>and</strong> use <strong>and</strong> cover changes<br />

as transition process from one state to other states. Typical examples are models based on<br />

cellular automata <strong>and</strong> Markov process models, such as the two models used in the present study.<br />

2. Material<br />

We used the programs DINAMICA EGO <strong>and</strong> L<strong>and</strong> <strong>Change</strong> Modeler in IDRISI for LUCC<br />

modeling. DINAMICA EGO is a cellular automata-based model which has been applied in a<br />

variety of studies, including modeling tropical deforestation (Soares-Filho et al. 2002 <strong>and</strong> 2006;<br />

Cuevas <strong>and</strong> Mas 2008) <strong>and</strong> urban growth <strong>and</strong> dynamics (Godoy <strong>and</strong> Soares-Filho 2008). L<strong>and</strong><br />

<strong>Change</strong> Modeler (available in IDRISI) provides tools for the assessment <strong>and</strong> projection of l<strong>and</strong><br />

cover change, <strong>and</strong> their implications for species habitat <strong>and</strong> biodiversity (Eastman 2006; Gontier<br />

et al. 2009; Koi <strong>and</strong> Murayama 2010). Statistical analysis <strong>and</strong> graphs were created using R (R<br />

Development Core Team 2009).<br />

* Corresponding author. Tel.: +52 443 328 38 35 - Fax: +52 443 38 80<br />

Email address: jfmas@ciga.unam.mx<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Assessing “spatially explicit” l<strong>and</strong> use/cover change models<br />

509<br />

Modeling was carried out using the data set supplied with the IDRISI tutorial which consists of<br />

l<strong>and</strong> cover (LC) maps <strong>and</strong> ancillary information from a rapidly changing area in the Bolivian<br />

lowl<strong>and</strong>s (Eastman 2009). The data used in the present study are the LC maps of 1986 <strong>and</strong> 1994<br />

<strong>and</strong> several maps used as explanatory variables (maps of distance from urban areas, distance<br />

from roads, slope, distance from disturbance, elevation).<br />

3. Methodology<br />

The present study aimed at: 1) creating modeled LC maps using DINAMICA <strong>and</strong> LCM; <strong>and</strong><br />

2) assessing these maps using two approaches, a) based on the spatial coincidence,, <strong>and</strong><br />

b) computing l<strong>and</strong>scape metrics.<br />

3.1. LUCC Modeling<br />

LUCC are modeled empirically by using past change to develop a mathematical model; <strong>and</strong> GIS<br />

data layers influence the transition potential. The simulation procedures can be sub-divided into<br />

the following basic steps:<br />

1) Calibration: The model is calibrated using a map of LUCC obtained through the comparison<br />

of LC maps at two different dates (1986 <strong>and</strong> 1994 in the present case). The quantity of each type<br />

of change is computed from a Markov matrix, which is the st<strong>and</strong>ard procedure in DINAMICA<br />

<strong>and</strong> LCM. A spatial analysis allows the identification of more likely change locations using a<br />

set of explanatory variables. Based upon the relationship between the different transitions <strong>and</strong><br />

the explanatory variables, maps of change potential are produced for each transition. In the<br />

present study, the DINAMICA model uses the map of probability elaborated by the LCM<br />

artificial neural network (ANN) in order to obtain comparable results.<br />

2) Simulation: A prospective LC map is created based upon the expected quantity of changes<br />

(Markov matrix). DINAMICA <strong>and</strong> IDRISI use a cellular automata approach in order to obtain a<br />

proximity effect <strong>and</strong> eventually simulate l<strong>and</strong>scape pattern. In IDRISI, the process involves a<br />

3x3 filter which reclassifies pixels to incorporate the effects of neighboring pixels on a current<br />

pixel value <strong>and</strong> there is no option to control the CA behavior. DINAMICA uses two<br />

complementary transition functions: 1) the Exp<strong>and</strong>er; <strong>and</strong> 2) the Patcher. The first process is<br />

dedicated only to the expansion or contraction of previous patches of a certain class. The second<br />

process generates new patches through a seeding mechanism. The user can set parameters to<br />

control the size <strong>and</strong> shape of the simulated patches, such as mean patch size, patch size<br />

variance, <strong>and</strong> isometry. Additionally, a “prune factor” allows simulated changes to occur in less<br />

likely areas.<br />

3.2. Model assessment<br />

The evaluation of the LC prospective map was based on the comparison between the simulated<br />

<strong>and</strong> the observed (true) map using two approaches: a) the spatial coincidence between modeled<br />

<strong>and</strong> true change; <strong>and</strong>, b) the spatial pattern of modeled <strong>and</strong> true change patches.<br />

In order to assess the spatial coincidence between simulated <strong>and</strong> true changes, we used the<br />

fuzzy similarity test based on the concept of fuzziness of location, in which a representation of a<br />

cell is influenced by the cell itself <strong>and</strong> by the cells in its neighborhood (Hagen 2003). Two-way<br />

comparison was conducted, applying the fuzziness to the simulated <strong>and</strong> the true maps of change<br />

in turn. As r<strong>and</strong>om maps tend to score higher, we picked up the minimum fit value from the<br />

two-way comparison. In order to assess the spatial configuration of simulated <strong>and</strong> true changes,<br />

we calculated, for each transition, the amount of change with respect to the map of change<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Assessing “spatially explicit” l<strong>and</strong> use/cover change models<br />

510<br />

probability. For this, a map of susceptibility categories was first obtained by reclassifying<br />

susceptibility maps into 10 categories <strong>and</strong> overlaying this with the maps of changes.<br />

Additionally, some metrics used to characterize l<strong>and</strong>scape were computed, such as the number<br />

<strong>and</strong> the size of the patches (mean <strong>and</strong> st<strong>and</strong>ard deviation) <strong>and</strong> total edge (mean <strong>and</strong> st<strong>and</strong>ard<br />

deviation). In the present study, as we are interested in assessing l<strong>and</strong>scape pattern simulation<br />

rather than predictive performance, we simulated a 1994 LC map from the model calibrated<br />

over the period 1986-94 (i.e. the simulation <strong>and</strong> calibration periods are the same).<br />

4. Result<br />

4.1. LUCC Modeling<br />

During 1986-94, the main LUCC transitions were the conversion to anthropogenic disturbance<br />

of:<br />

Transition 1:Deciduous mature forest.<br />

Transition 2: Savanna.<br />

Transition 3: Amazonian mature forest.<br />

Transition 4: Woodl<strong>and</strong> savanna.<br />

Only these 4 principal transitions were modeled using as explanatory variables the distance<br />

from 1986 urban areas, the distance from roads, the slope, the distance from 1986 disturbance,<br />

the elevation <strong>and</strong> the 1986 LC map. The two programs were used to build 1994 simulated LC<br />

maps (Figure 1). In the case of DINAMICA, various settings of prune factors, patch sizes <strong>and</strong><br />

isometry were tested.<br />

True 1994 LC map<br />

Simulated 1994 LC map (DINAMICA)<br />

Simulated 1994 LC map (LCM)<br />

Figure 1: True 1994 LC map <strong>and</strong> modeled maps by DINAMICA <strong>and</strong> LCM (Zoom on the Southeastern<br />

part of the study area, only the category anthropogenic disturbance is represented)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Assessing “spatially explicit” l<strong>and</strong> use/cover change models<br />

511<br />

4.2. Model assessment<br />

The fuzzy similarity index indicates that coincidence between the true changes <strong>and</strong> the changes<br />

modeled by LCM is much higher than with DINAMICA using little tolerance (fuzzy tolerance<br />

distance < 1000 m). This result was expected as LCM tends to collocate simulated changes only<br />

in the areas with higher change potential. Since DINAMICA makes an attempt to create patches<br />

<strong>and</strong> simulates change in less likely areas (if the prune factor value is set high), the coincidence<br />

between the true <strong>and</strong> simulated changes is likely to be lower. However, with higher fuzzy<br />

tolerance values, DINAMICA presents a higher score because it has some (fuzzy) coincidence<br />

of simulated patches located in less likely areas. This does not occur with LCM that restricts the<br />

simulated change to the more susceptible areas only. Therefore, DINAMICA presents a better<br />

coincidence “as a broad picture” whereas LCM exhibits a better coincidence on a per-pixel<br />

comparison or with little fuzzy tolerance.<br />

Table 1 shows that with DINAMICA it was possible to obtain for each transition simulated<br />

patches of change that present broadly the same size as true patches of change. In the case of<br />

LCM, the simulation produced some very large patches corresponding to the higher<br />

susceptibility areas, resulting in larger values for the mean <strong>and</strong> the st<strong>and</strong>ard deviation of patch<br />

size. A similar pattern can be observed for patch edge lengths. However, there are fewer patches<br />

on the true change map than on either of the simulated change maps. The map modeled with<br />

LCM has 47% fewer patches than the true map.<br />

Table 1: L<strong>and</strong>scape metrics for true <strong>and</strong> simulated maps<br />

Metric Transition True <strong>Change</strong>s DINAMICA LCM<br />

Number of patches Transition 1 332 204 192<br />

Transition 2 326 208 190<br />

Transition 3 666 615 299<br />

Transition 4 1017 804 547<br />

Patch size (mean / Transition 1 5.7 / 6.9 9.3 / 9.3 9.7 / 36.2<br />

st<strong>and</strong>ard deviation)<br />

Transition 2 9.8 / 18.4 15.4 / 21.3<br />

16.7 / 41.7<br />

Transition 3 17.1 / 46.3 18.5 / 23.5 36.1 / 162.1<br />

Transition 4 19.4 / 45.6<br />

35.4 / 151.6<br />

24.5 / 32.4<br />

Patch edge length Transition 1 1058.1 / 802.4 1551.5 / 1262.8 1417.2 / 2936.7<br />

(mean / st<strong>and</strong>ard<br />

Transition 2<br />

2097.1 / 2056.8 2133.2 / 3690.2<br />

deviation)<br />

1465 / 1835.6<br />

Transition 3 1856.8 / 2604.1 2245.4 / 2240.3 3107.9 / 9325.7<br />

Transition 4 2233.2 / 3343.9 2889 / 3184.5 3240.2 / 9436.7<br />

Figure 2 shows that true changes do not occur only in more susceptible areas <strong>and</strong> that this<br />

tendency depends on the transition. For example, transition 3 occurred mainly in the more<br />

susceptible area whereas transition 2 is frequent even in areas with medium susceptibility. The<br />

changes simulated by LCM are limited to the areas with higher susceptibility. The setting of the<br />

prune factor allowed DINAMICA to generate a map with a distribution of change closer to the<br />

observed change.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Assessing “spatially explicit” l<strong>and</strong> use/cover change models<br />

512<br />

Figure 2. Distribution of change in categories of change susceptibility.<br />

5. Discussion<br />

DINAMICA was able to generate more realistic prospective LC maps with respect to l<strong>and</strong>scape<br />

pattern because it provides parameters to control the CA behavior. However, the best manner of<br />

producing a prospective LC map, where simulated changes fit better with the true changes, is by<br />

thresholding the susceptibility map, because the majority of the changes occur in the more likely<br />

locations. The realism of the l<strong>and</strong>scape pattern in the prospective LC map is obtained at the<br />

expense of the accuracy of the locations of the change. This is particularly obvious when<br />

models simulate the occurrence of changes in unlikely areas. The prune factor in DINAMICA<br />

also allows the occurring of change in less likely areas. When modeling aims at producing LC<br />

maps that can represent a possible future given a certain scenario, the accuracy of the spatial<br />

allocation of change is not necessarily a critical issue. For example, in the assessment of LUCC<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.-F. Mas et al. 2010. Assessing “spatially explicit” l<strong>and</strong> use/cover change models<br />

513<br />

on biodiversity, it can be important to know that homogeneous forest areas will be perforated by<br />

small agriculture fields although the exact location of the fields remain unknown.<br />

However, the common procedures of assessment of a prospective LC map are based upon the<br />

spatial coincidence of the simulated map <strong>and</strong> a “true” observed map. Therefore, the modeling<br />

<strong>and</strong> the assessment procedures have to be adapted to the critical feature the model has to<br />

achieve. When l<strong>and</strong>scape pattern is an important feature, the computing of l<strong>and</strong>scape metrics<br />

can provide valuable insights to evaluate the model performance.<br />

6. Acknowledgements<br />

This research has been supported by the Consejo Nacional de Ciencia y Tecnología -<br />

CONACyT (Sabbatical <strong>and</strong> posdoctoral stays at the University of California - Santa Barbara of<br />

the first <strong>and</strong> second author respectively) <strong>and</strong> the Dirección General de Asuntos del Personal<br />

Académico (DGAPA) at the Universidad National Autónoma de México (additional support for<br />

the sabbatical stay)<br />

References<br />

Cuevas, G <strong>and</strong> Mas, J-.F., 2008. L<strong>and</strong> use scenarios: a communication tool with local<br />

Communities. In: M. Paegelow <strong>and</strong> M.T. Camacho (Eds). Modelling Environmental<br />

Dynamics, Springer-Verlag. pp. 223-246.<br />

de Sherbinin, A., 2002. CIESIN Thematic Guide to L<strong>and</strong> L<strong>and</strong>-Use <strong>and</strong> L<strong>and</strong> L<strong>and</strong>-Cover<br />

<strong>Change</strong> (LUCC). Center for International Earth Science Information Network (CIESIN)<br />

Columbia University Palisades, NY, USA.<br />

http://sedac.ciesin.columbia.edu/guides/lu/CIESIN_LUCC_TG.pdf.<br />

Eastman, J.R. 2009. IDRISI taiga Tutorial. Accessed in IDRISI 16.05. Worcester, MA: Clark<br />

University: 333 p.<br />

Eastman, J.R., Van Fossen, M.E. <strong>and</strong> Solarzano, L.A., 2005. Transition Potential Modeling for<br />

L<strong>and</strong> Cover <strong>Change</strong>, in: D. Maguire, M. Goodchild <strong>and</strong> M. Batty (Eds). GIS, Spatial<br />

Analysis <strong>and</strong> Modeling,., ESRI Press Redl<strong>and</strong>s, California.<br />

Godoy, M.M.G. <strong>and</strong> Soares-Filho, B.S., 2008. Modelling intra-urban dynamics in the Savassi<br />

neighbourhood, Belo Horizonte city, Brazil. In: M. Paegelow <strong>and</strong> M.T. Camacho (Eds).<br />

Modelling Environmental Dynamics. Springer-Verlag, pp. 319-338.<br />

Gontier, M., Mörtberg, U. <strong>and</strong> Balfors, B. in press. Comparing GIS-based habitat models for<br />

applications in EIA <strong>and</strong> SEA. Environmental Impact Assessment Review.<br />

Hagen, A., 2003. Fuzzy set approach to assessing similarity of categorical maps. International<br />

Journal of Geographical Information Science, 17(3): 235–249.<br />

Koi, D.D. <strong>and</strong> Murayama, Y., 2010, Forecasting Areas Vulnerable to <strong>Forest</strong> Conversion in the<br />

Tam Dao National Park Region, Vietnam. Remote Sensing, 2, 1249-1272.<br />

R Development Core Team, 2009. R: A language <strong>and</strong> environment for statistical computing. R<br />

Foundation for Statistical Computing,Vienna, Austria, URL http://www.R-project.org.<br />

Soares-Filho, B.S., Pennachin, C.L. <strong>and</strong> Cerqueira, G., 2002. DINAMICA – a stochastic cellular<br />

automata model designed to simulate the l<strong>and</strong>scape dynamics in an Amazonian<br />

colonization frontier. Ecological Modelling, 154(3):217-235.<br />

Soares-Filho, B.S., Nepstad, D., Curran, L., Voll, E., Cerqueira, G., García, R.A., Ramos,<br />

C.A., Mcdonald, A., Lefebvre, P. <strong>and</strong> Schlesinger, P., 2006. Modelling conservation in<br />

the Amazon basin. Nature, 440:520-523.<br />

Veldkamp, A. <strong>and</strong> Lambin, E., 2001. Predicting l<strong>and</strong>-use change. Agriculture, Ecosystems <strong>and</strong><br />

Environment, 85:1-6.<br />

Xiang, W-N <strong>and</strong> Clarke, K.C., 2003. The use of scenarios in l<strong>and</strong>-use planning. Environment<br />

<strong>and</strong> Planning B: Planning <strong>and</strong> Design, 30(6) 885-909.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Matos et al. 2010. Economic valuation of environmental goods <strong>and</strong> services<br />

514<br />

Economic valuation of environmental goods <strong>and</strong> services<br />

Alda Matos 1* , Paula Cabo 2 , Isabel Ribeiro 2 & António Fern<strong>and</strong>es 2<br />

1<br />

Polytechnic Institute of Bragança, Portugal<br />

2<br />

CIMO - Mountain Research Centre, Portugal<br />

Abstract<br />

The economic valuation of environmental goods <strong>and</strong> services (EVEG&S) results of the<br />

increasing concern with the quality of industrial products <strong>and</strong> the reduction of social welfare.<br />

The EVEG&S presents the direct <strong>and</strong> indirect costs <strong>and</strong> benefits of quantitative <strong>and</strong> qualitative<br />

environmental changes in goods <strong>and</strong> services <strong>and</strong> corresponding impacts. This is particularly<br />

important in the valuation of investment projects <strong>and</strong> governmental policies. This study consists<br />

in a survey of environmental appraisal methods, focusing into the hypothetical <strong>and</strong><br />

complementary market based ones. The review reveals that evaluation of environmental quality<br />

is very complex. In fact, for each criterion there are several assumptions that are inapplicable to<br />

all situations. Effectively, despite the evident complementarity of conventional goods<br />

environmental quality, the values attributed to these resources could be underestimated <strong>and</strong><br />

complementary <strong>and</strong> substitute markets can be inefficient parameters.<br />

Keywords: Economic valuation, natural resources, markets, environmental goods<br />

1. Introduction<br />

The economic growth is based on wealth creation, based on a process of dominance <strong>and</strong><br />

transformation of the Nature. The modern society is guilty of wild exploitation of natural<br />

resources, neglecting the damages of productive activities. The dem<strong>and</strong> <strong>and</strong> improper use of the<br />

natural resources increases daily. With this speedy environmental harm, environmental<br />

protection st<strong>and</strong>s out as one of the current <strong>and</strong> future major challenges for humanity. The<br />

economic appraisal of environment results of the increasing concern with protection <strong>and</strong><br />

preservation of natural resources <strong>and</strong> consumers’ requests for quality industrial products,<br />

simultaneous with the reduction of social welfare, as consequence of the quality <strong>and</strong> amount of<br />

these resources. The economic appraisal emerges as a measuring tool of environmental goods<br />

<strong>and</strong> services <strong>and</strong> of the impacts of environmental degradation <strong>and</strong> depletion, determining the<br />

direct <strong>and</strong> indirect costs <strong>and</strong> benefits of qualitative <strong>and</strong> quantitative changes. It is gathering<br />

importance in the evaluation of investment projects, governmental policies <strong>and</strong> international<br />

trade. This paper focuses on this problematic. The paper consists of a critical analysis of the<br />

economic appraisal criteria of environmental goods <strong>and</strong> services. Particularly of the methods<br />

that make use of hypothetical <strong>and</strong> complementary market goods.<br />

2. Economic Valuation of Environmental Goods <strong>and</strong> Services<br />

Based on the externality notion, Foladori (1997) defends that negative trends inherent to free<br />

market can be beated through environmental appraisal with the inclusion of prices in economic<br />

analysis, via policies that attenuate environmental problems. Schweitzer (1990) beliefs that<br />

environmental appraisal is fundamental to prevent the depletion of natural resources.<br />

The environmental appraisal emerges as a set of techniques <strong>and</strong> methods to quantify the<br />

expectations of benefits/costs derived from the use of environmental assets, carrying out<br />

* Corresponding author. Tel.: (+351) 273 303 242 - Fax: (+351) 273 325 405<br />

Email address: alda@ipb.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Matos et al. 2010. Economic valuation of environmental goods <strong>and</strong> services<br />

515<br />

benefittings or infliction of environmental damages. The economic value of an environmental<br />

good consists of the estimate of a monetary value for this good, in opposition to other available<br />

goods. However, some times, it’s difficult to aggregate all the effects in a single indicator.<br />

The economic value of environmental resources (EVER) results from its attributes, <strong>and</strong> these<br />

can be associated to the use (direct, indirect <strong>and</strong> option) or non-use of the resource, i.e., its<br />

simple existence. EVER purposes a fee for environmental resources’ use <strong>and</strong>/or preservation.<br />

The genesis is the protection of current <strong>and</strong> future generations’ interests. Thus, use value (UV)<br />

is the value attributed by people who use or usufruct of the environmental good to satisfy their<br />

needs. The non use value (NUV) is dissociate of the use because it derives from a moral,<br />

cultural, ethical or altruistic position regarding the rights of existence of other living species or<br />

the preservation of natural assets although that do not represent current or future use for them.<br />

While slightly different classifications exist, they result the same. Still, controversy subsists<br />

regarding existence (EV) <strong>and</strong> option (OV) values, since the EV represents the individual will to<br />

preserve a set of environmental resources for future generations’ direct <strong>and</strong>/or indirect use. Thus,<br />

the conceptual question is if a value defined like so is closer associated with the OV or the EV.<br />

Equally, the legacy value (in this definition mixed with the EV) can be independent (Figure1).<br />

However, for EVER matters that the individuals point out the most trustworthy values possible,<br />

independently of the current or future use.<br />

Economic Value of Environmental Resources<br />

Non Use Value<br />

Direct Use Value<br />

Indirect Use<br />

Value<br />

Option Value Existence Value Legacy Value<br />

Figure 1: Different economic values of environmental resources<br />

The environmental appraisal difficulty increases inversely as function of the resources’ use. The<br />

choice of the criterion depends on the knowledge of the ecological dynamics of the study object,<br />

the purpose of the valuation, the availability of information <strong>and</strong> the hypotheses adopted.<br />

Environmental economics classifies the valuation techniques in production function methods –<br />

marginal productivity method <strong>and</strong> markets of substitute goods method – <strong>and</strong> dem<strong>and</strong> function<br />

methods – methods that utilize markets of complementary goods (hedonic prices <strong>and</strong> travel<br />

costs methods) <strong>and</strong> hypothetical markets (method of contingent valuation). May <strong>and</strong> Motta<br />

(1994) refer that production function methods analyze environmental resources associated to the<br />

production of a private good <strong>and</strong>, generally, assume that supply variations do not influence<br />

market prices. The dem<strong>and</strong> function methods admit that changes in resource availability modify<br />

individual wellbeing <strong>and</strong>, therefore, it’s possible to identify individual measures of Willingness<br />

to Pay (WTP) or Willingness to Accept (WTA) regarding to these variations. These are the<br />

methods under this study review.<br />

2.1. Analysis of the Dem<strong>and</strong> Function Methods<br />

To Dixon et al. (2001), subjective valuation methodologies can assess consumers revealed or<br />

expressed preferences, in real or fictitious markets, relating it with individuals’ utility functions.<br />

Mainly, these methodologies use substitute market prices or contingent values (Table 1)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Matos et al. 2010. Economic valuation of environmental goods <strong>and</strong> services<br />

516<br />

Table 1: Subjective valuation criteria<br />

Complementary Goods Market<br />

Hedonic Prices<br />

Property value Revealed behaviour Substitute market prices<br />

Wages differential Revealed behaviour Substitute market prices<br />

Travel Costs Revealed behaviour Substitute market prices<br />

Hypothetical Markets<br />

Contingent Valuation Expressed behaviour Contingent value<br />

Revealed preferences analysis is based on real markets of goods <strong>and</strong> services (MG&S) affected<br />

by environmental impact, in which folks pick between levels of environmental quality <strong>and</strong> other<br />

goods. When resources are not marketly traded, the economic analysis aims to estimate the<br />

economic value as if the market exists. The analysis of expressed behaviours is used when is not<br />

possible to valuate the environmental impacts, even dissimulatly, through real markets.<br />

2.1.1. Markets of Complementary Goods<br />

The urban amenities are not restricted to natural features, as green areas, beaches <strong>and</strong> climate.<br />

The concept also adds the goods (or evils) created by men, as traffic, pollution, recreation areas<br />

<strong>and</strong> safety. The study of these environmental attributes permits to underst<strong>and</strong> the impact (<strong>and</strong><br />

the changes) of the cities physical space on their inhabitants’ wellbeing, as well as the effect on<br />

the real estate market value. To quantify urban amenities is not a simple task. Although the real<br />

estate market has supply, dem<strong>and</strong> <strong>and</strong> an equilibrium price, it’s not possible to visualize the<br />

market prices of environmental amenities, since trade of oxygen, l<strong>and</strong>scape, recreation areas, or<br />

traffic, pollution <strong>and</strong> noise, does not exist.<br />

The hedonic prices (MHP) <strong>and</strong> travel costs (TCM) models are the more adjusted criteria to<br />

decode this information. They are based in the preferences revealed by consumers in a substitute<br />

market, <strong>and</strong> use it to assess individuals’ wellbeing, in view of environmental quality changes.<br />

The MHP has been widely used in the real estate market to measure the marginal value of<br />

natural or structural attributes, <strong>and</strong> estimate the correlated social-environmental variables. This<br />

method is based on the recognition of the complementary attributes of a specific private<br />

composed good to environmental goods or services (Motta, 1998). This complementarity<br />

discloses the price of the environmental attribute implicit in the market price.<br />

The method of property value (<strong>and</strong> wages differential method) solely valuates UV. It only looks<br />

into the appraisal of environmental functions/services that directly affect the market prices of<br />

related goods. The MHP considers a heterogeneous good as a closed package, with specific<br />

attributes, where the marginal price of each one is estimated, based on the analysis of the good<br />

observed value <strong>and</strong> attributes’ respective amounts (Rosen, 1974). It presumes that families,<br />

when look for housing, are worried about what exists inside <strong>and</strong> outside (the amenities) of the<br />

property. These amenities (distance to workplace, proximity of parks, beach, schools, quality of<br />

air, water, sonorous pollution, l<strong>and</strong>scape, etc.) will imply variations in the asset usufruct.<br />

Similarly, the price of a specific l<strong>and</strong> does not depend solely on its patrimonial value, but also<br />

on the actual value of the net benefits generated by soil productivity over time. Thus, because<br />

productivity levels differ, different l<strong>and</strong> fractions have different price levels. Additionally, the<br />

environmental features, as air quality, water disposal (for irrigation) or erosion, affect the soil<br />

quality for agriculture, <strong>and</strong> thus, its price. The MHP estimates the quantitative differences of the<br />

attributes, using market prices of goods or costs of services essentials in the formation of these<br />

prices/costs. These discrepancies are valued by individuals, reflecting their WTP when the<br />

environmental attributes vary. According to this paradigm, two houses with identical physical<br />

attributes, situated in different ecological <strong>and</strong> social contexts, will have different prices.<br />

The MHP catches only the use values. EV is not catched because of the weak complementarity.<br />

When the dem<strong>and</strong> for a specific environmental attribute is null, the dem<strong>and</strong> for the composed<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Matos et al. 2010. Economic valuation of environmental goods <strong>and</strong> services<br />

517<br />

good is also null. Redondo (1999) refers that MHP reproduces the changes in the UV of a<br />

specific place inhabitants, but are merely informative in the “passer-bys” case (individuals that<br />

do not have fixed residence there, but sporadically travel to it) <strong>and</strong> reveal nothing about NUV<br />

(associated with individuals that do not use the place). The adoption of this criterion requires the<br />

existence of a reasonable mobility in real state market, so that folks can reveal their WTP in an<br />

environmental context where is possible to choose between houses with different attributes <strong>and</strong><br />

without prohibitive transaction costs (costs of house search <strong>and</strong> changing, taxes <strong>and</strong> duties<br />

associated with the sale/purchase). The improve of life quality conditions in a specific<br />

neighbourhood, does not necessarily imply a change in housing prices, due to the people weak<br />

mobility to try another neighbourhood <strong>and</strong> their WTP for usufruct of it.<br />

Dixon et al. (1994), Comune et al. (1995), Motta (1998), Redondo (1999), among others,<br />

emphasize that MHP efficient results require the assembling of detailed <strong>and</strong> faithful information<br />

on the characteristics of the asset under valuation. An exhaustive survey on the environmental<br />

indicators is crucial. Namely, the attributes that influence the assets price: property features<br />

(size, degree of conservation, benefittings); commercial, transport <strong>and</strong> education services; local<br />

community quality (neighbouring, criminality) <strong>and</strong> social-economic information of a sample<br />

representative of local owners. Additionally, the attribute under valuation must be clearly <strong>and</strong><br />

precisely defined, in order to successfully isolate it from the subject other attributes. The same<br />

authors refer the operative difficulties in the econometrical estimation of the hedonic functions<br />

due to the omission of relevant variables, attributes’ multicollinearity, <strong>and</strong> functional form<br />

identification, among others. Additionally, property prices can be underestimated due to minor<br />

property tax transfer or to attenuate the effect of patrimonial variations. The alternative would<br />

be to use leasing values. So, this method must only be used in case of high correlation between<br />

property price <strong>and</strong> environmental attribute, when is possible to catch all the attributes<br />

influencing the real estate market equilibrium price <strong>and</strong> when the hypotheses adopted for the<br />

estimate of the consumer surplus are realistic.<br />

The TCM is used in the valuation of environmental resources as parks <strong>and</strong> recreational sites, but<br />

also to quantify externalities of urban collective transport projects. The TCM basic premise is<br />

that the costs of accessing to a place directly influence its visits number. This method associates<br />

the environmental resources value to its recreation value. The benefits of a specific investment<br />

are quantified in function of the (estimated costs by the) curve of dem<strong>and</strong> of the activity, based<br />

on the study of their users’ expenditures (in time <strong>and</strong> transportation costs).<br />

The TCM is based on a preferences approach. The individual reveals his choices buying specific<br />

goods associated with the use of an environmental good. This approach requires interviews to<br />

the visitors in order to determine their st<strong>and</strong>ard use <strong>and</strong> to gather information on the number of<br />

visitors; visitors’ geographic, social <strong>and</strong> economical characteristics; motive, duration <strong>and</strong><br />

frequency of the visit; transportation mean <strong>and</strong> costs associated to the trip... The data collected<br />

will be used to estimate a visitation rate by region of origin, the total travel costs <strong>and</strong> link it with<br />

the visits frequency, establishing a dem<strong>and</strong> correspondence. Each individual income matches a<br />

dem<strong>and</strong> function, given that each person is WTP a determined price in exchange for an amount<br />

of the product. The curve of dem<strong>and</strong> for visit for each region <strong>and</strong> the aggregated dem<strong>and</strong> curve<br />

are determined. Then, visitation dem<strong>and</strong> function is used to estimate the consumer surplus,<br />

which represents the economic value of the recreational site.<br />

The disadvantages of the TCM application are related with the individual visits’ duration, the<br />

possibility of resources deterioration, the distance (it’s expected that distant residents visit less<br />

the recreational site, while they can actually have longer visits), the difficulty in the exclusion of<br />

services not associated to the site (multiple trip objectives <strong>and</strong> destinations), the merely capture<br />

the visits direct <strong>and</strong> indirect UV <strong>and</strong> the monetary value of the time spent by the visitor<br />

(overvaluing the recreation cost, due to price distortions in the labour market). Other<br />

disadvantages are related with the premises assumed in the estimation of the curve of dem<strong>and</strong>;<br />

the need of reliable data; high application costs; dependency of statistical methods; <strong>and</strong> the no<br />

consideration of NUV components.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Matos et al. 2010. Economic valuation of environmental goods <strong>and</strong> services<br />

518<br />

2.1.2. Hypothetical markets<br />

To Hicks (1939), the estimation of a change in consumer wellbeing can be carried out by its<br />

income variation, introducing two measures of value that support the economic valuation of<br />

environmental impacts. The measures are compensatory <strong>and</strong> equivalent variations <strong>and</strong> are<br />

linked with variations in consumers’ utility <strong>and</strong> preferences (WTP <strong>and</strong> WTA).<br />

According to Comune et al. (1995), the Method of Contingent Valuation (MCV) aggregates a<br />

set of techniques used in research to estimate the EVEG&S, based on consumers’ preferences.<br />

These techniques are based on individual budgetary evaluations, given an increase or decrease<br />

in the quality or amount of an environmental good or service, in a hypothetical scenario. This is<br />

the only method that allows valuating the UV <strong>and</strong> NUV of environmental resources. Its domain<br />

of application is the valuation of wildlife, protection of habitats <strong>and</strong> measurement of the UV of<br />

leisure <strong>and</strong> recreational sites. According to Dixon et al. (1994), the MCV constitutes the only<br />

alternative to attain economic value estimates when in presence of distortions in environmental<br />

MG&S, there are no effective market nor substitute markets for it. Walnut et al. (2000) explain<br />

that the theoretical concept of MCV is consumer theory (consumer choice <strong>and</strong> consumer<br />

surplus). The individual WTP discloses, through the graduation of the marginal utility, the best<br />

estimate of its dem<strong>and</strong> scale, <strong>and</strong> thus, quantifying social welfare measures. The consumer<br />

choices are based on the utility maximization premise, under budgetary restriction. The<br />

consumer surplus valuates the different degrees of individuals’ preferences for various goods<br />

<strong>and</strong> services revealed when consumers go to the market <strong>and</strong> pay a specific amount for them.<br />

The MCV uses questionnaire techniques to valuate consumers expressed preferences, <strong>and</strong><br />

clearly describe the good to quantify. In order to the respondents declare <strong>and</strong> quantify their real<br />

preferences, this method simulates scenarios with characteristics analogous to the existing in the<br />

real world. Based on personal opinions, constructs a hypothetical market <strong>and</strong> quantify WTP<br />

(payment for a wellbeing improvement) <strong>and</strong> WTA (reimbursement for a wellbeing loss)<br />

according to variations in the availability of environmental resources. The intended result is to<br />

reach maximum WTP for a given benefit, the minimum compensation to abdicate of the benefit<br />

or WTA for environmental damage. Finally, average WTP/WTA is calculated, the populations<br />

are added <strong>and</strong> thus obtaining the estimates of the value attributed to the environmental good.<br />

Comune et al. (1995: 64) point out that one of the advantages of this type of methodology<br />

consists, precisely, of producing estimates of values that could not be obtained by other ways.<br />

To Macedo (2002), the limitations of these methods derive from individuals apparently<br />

contradictory behaviours, according with the roles adopted in face of the environmental good.<br />

The author refer that most of the folks is propense to establish extremely high values to admit<br />

the loss of a natural resources <strong>and</strong> excessively low values in the hypothesis of having to pay to<br />

assure the it protection.<br />

The MCV can bear ambiguous results due to bias, resulting from the market fictitious feature<br />

<strong>and</strong> from quality of the individuals’ information. The respondents can not reveal the real WTP<br />

or WTA due to their reduced experience, mostly for the WTA case. Moreover, the interviewer<br />

can induce answers. And, having no commitment with an effective payment, the vehicle used<br />

can affect the result.<br />

3. Final remarks<br />

This paper describes several methodologies of evaluation of environmental goods, showing that<br />

the quantification of environmental quality is not simple to get. None of the different existing<br />

methods adjust to all situations. Each criterion is limited to specific conditions, <strong>and</strong> therefore<br />

unacceptable <strong>and</strong> inapplicable in others.<br />

The economic indicators are precious tools for unique decision making, however, as society<br />

general knowledge of ecosystem functions is reduced, they become limited, consequently<br />

overvaluing individual preferences, that is, overvalue a subsystem in detriment of another<br />

possibly more valuable for the project. Therefore, the EVEG&S incurs in an implicit<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Matos et al. 2010. Economic valuation of environmental goods <strong>and</strong> services<br />

519<br />

subjectivity of the importance of the scale <strong>and</strong> the definition of object of study. With the<br />

existing subjectivity, we can argue about the multiplicity of the value, since different exercises<br />

of quantification origin distinct results, according to the purpose <strong>and</strong> the methodology applied.<br />

Such multiplicity does not reduce the importance of the valuation as an analysis technique, but,<br />

each result can be influenced by different perspectives, <strong>and</strong>, thus alerting to the values’ partiality.<br />

The EVEG&S should be executed in partnership by environmental, social scientists <strong>and</strong><br />

economists, with inter, trans <strong>and</strong> multidisciplinary dimension to avoid the risk of<br />

indiscriminately choose methods inadequate to the reality in study. The EVEG&S is of extreme<br />

utility for the decision making, but has limits of scientific uncertainty that go beyond economic<br />

science. Therefore, it would be of all interest a bigger scientific cooperation in this area, in order<br />

to increase quality to the current state of the art, since the resources’ values can be<br />

underestimated <strong>and</strong> the complementary or substitute markets can be inefficient parameters.<br />

References<br />

Comune, A.; Grasso. M.; Tognella, M. <strong>and</strong> Schaeffer, Y.,1995.Aplicação de Técnicas de<br />

Avaliação Econômica ao Ecossistema Manguezal. Valor<strong>and</strong>o a Natureza.<br />

Dixon, J.; Scura, L.; Carpenter, R. <strong>and</strong> Sherman, P.,1994. Economic Analysis of Environmental<br />

Impacts. Earthscan Publications Ltd. London.<br />

Foladori, G., 1997. A Economia Frente à Crise Ambiental. Revista de Economia, 23(21).<br />

Hicks, J., 1939. The Foundations of Welfare Economics. Economic Journal, 49:696-712.<br />

Loomis, J. <strong>and</strong> González-Cabán, A., 1998. A Willingness to Pay for Protecting Acres of Spotted<br />

Owl Habitat from Fire. Ecological Economics, 25:315-322.<br />

Loomis, J.; González-Cabán, A. <strong>and</strong> Gregory, R., 1996. A Contingent Valuation Study of the<br />

Value of Reducing Fire Hazards to Old-growth <strong>Forest</strong>s in the Pacific Northwest. Pacific<br />

Southwest Research Station. USDA, <strong>Forest</strong> Serv. Berkeley. CA.<br />

Macedo, Z., 2002. Os Limites da Economia na Gestão Ambiental. Margem, 15:203-222.<br />

May, P. <strong>and</strong> Motta, R., 1994. Valor<strong>and</strong>o a Natureza: Análise Económica para o<br />

Desenvolvimento Sustentável. Editora Campus. Rio de Janeiro.<br />

Motta, R., 1998. Manual para Valoração Económica de Recursos Ambientais. Ministério do<br />

Meio Ambiente, dos Recursos Hídricos e da Amazônia Legal, Brasília, 218p.<br />

Moura, L., 2000. Economia Ambiental: Gestão de Custos e Investimentos. Ed. Juarez de<br />

Oliveira, S. Paulo, 200p.<br />

Munasinghe, M., 1992. Environmental Economics <strong>and</strong> Valuation of Development Decisions.<br />

Banco Mundial.<br />

Nogueira, J.; Medeiros, M. <strong>and</strong> Arruda, F., 2000. Valoração Econômica do Meio Ambiente:<br />

Ciência ou Empirismo Cadernos de Ciência e Tecnologia, 17(2):81-115.<br />

Pearce, D. <strong>and</strong> Turner, R., 1990. Economics of Natural Resources <strong>and</strong> the Environment.<br />

Harviester Wcatsheag, The Johns Hopkins University, Baltimore.<br />

Redondo, O., 1999. Entre la Economia y la Naturaleza. La Controversia sobre la Valoración<br />

Monetaria del Medio Ambiente y la Sustentabilidad del Sistema Económico. Los Libros<br />

de la Catarata, Madrid.<br />

Rosen, S., 1974. Hedonic Price <strong>and</strong> Implicit Markets: Product Differentiation in Pure<br />

Competition. Journal of Political Economics, 82:34-55.<br />

Santos, R.; Martinho, S. <strong>and</strong> Antunes, P., 2001. Avaliação Económica dos Impactes Ambientais<br />

do Sector Eléctrico. Estudo Sobre o Sector Eléctrico e Ambiente. 2º Relatório. Centro de<br />

Economia Ecológica e Gestão do Ambiente, Universidade Nova de Lisboa.<br />

Schweitzer, J., 1990. Economics, Conservation <strong>and</strong> Development: a Perspective from USAID.<br />

In: Vicent, J.; Crawfor, E. <strong>and</strong> Hoehn, J. (eds). Valuing Environmental Benefits in<br />

Developing Countries: Proceedings. East Lansing, Michigan State University.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Migliozzi et al. 2010. L<strong>and</strong>-use management <strong>and</strong> changes in Campania Region (Southern Italy)<br />

520<br />

L<strong>and</strong>-use management <strong>and</strong> changes in Campania Region (Southern<br />

Italy): examples from ten regional State <strong>Forest</strong>s<br />

A. Migliozzi 1 , F. Cona 1 , A. di Gennaro 2 , A. Mingo 1 , A. Saracino 1 & S. Mazzoleni 1<br />

1<br />

Department ARBOPAVE, University of Napoli “Federico II”, Via Università 100,<br />

80055 Portici (NA), Italy<br />

2<br />

Risorsa s.r.l., Via Raffaello 15, 80129 Napoli, Italy<br />

Abstract<br />

In the last 50 years the Italian territory has undergone a remarkable transformation as a<br />

consequence of anthropogenic activity . In the Campania Region, the main causes of change are<br />

urban expansion, ab<strong>and</strong>onment of mountainous <strong>and</strong> marginal agricultural areas <strong>and</strong> expansion<br />

of industrial settlements. We document l<strong>and</strong> use variation during the last 50 years <strong>and</strong> the<br />

changes occurred in ten Regional <strong>Forest</strong>s which contain mainly natural l<strong>and</strong> use patches<br />

The main problems of l<strong>and</strong> <strong>and</strong> forest management are discussed in the light of actions<br />

promoted by the promulgation of two important legislative instruments, the “Piano Territoriale<br />

Regionale (PTR)” <strong>and</strong> the “Piano <strong>Forest</strong>ale Generale (PFG)”, for the socio-economic<br />

revitalization of rural areas <strong>and</strong> for the sustainable management of agro-forestry systems.<br />

Keywords: Systems of L<strong>and</strong>s, Urban Sprawl, <strong>Forest</strong>ry, GIS<br />

1. Introduction<br />

In the last fifty years the agro-forestry l<strong>and</strong>scape of the Campania Region (Southern Italy) has<br />

undergone a remarkable transformation under the pressure of two opposing forces: urban sprawl<br />

(almost fivefold) <strong>and</strong> intensification <strong>and</strong> specialisation of agriculture in the coastal plains <strong>and</strong><br />

hills, contemporary to l<strong>and</strong> ab<strong>and</strong>onment in the mountain areas <strong>and</strong> in the inl<strong>and</strong> hills (Di<br />

Gennaro <strong>and</strong> Innamorato, 2005, Mazzoleni et al. 2004, Motti et al.2004, Migliozzi et al. 2001).<br />

The key areas of major l<strong>and</strong>scape transformation are situated along the sea coast, where the<br />

rapidly increasing urbanisation eroded the cultivated l<strong>and</strong> areas <strong>and</strong> the residual plain forests.<br />

The consequent congestion of the flat areas is a completely unbalanced use of soil, water <strong>and</strong><br />

l<strong>and</strong> resources. This is highlighted by the centrifugal expansion of urban areas at the expense of<br />

cultivated fields <strong>and</strong> natural vegetation <strong>and</strong> the contemporary centripetal recolonisation by<br />

wildl<strong>and</strong> vegetation of the ab<strong>and</strong>oned cultivated l<strong>and</strong>.. The soil loses its primary function <strong>and</strong><br />

becomes a free building space. The areas recolonised by natural vegetation, tends to interface<br />

directly with the exp<strong>and</strong>ing urban areas. Therefore water supply, pollution, management <strong>and</strong><br />

transport of solid <strong>and</strong> liquid waste increase. Furthermore natural vegetation dynamics on the<br />

ab<strong>and</strong>oned crops determines a homogenisation of the l<strong>and</strong>scape <strong>and</strong> modifies of the<br />

hydrological processes, leading to an increase of the hazard management issues (fire <strong>and</strong><br />

l<strong>and</strong>slide) at the urban-forest interface (Mazzoleni et al. 2004). The epochal dimension of this<br />

change, is the following: from 1960 to 2000 the urban space has increased from 20,000 ha to<br />

about 93,000 ha <strong>and</strong> the LUW ( L<strong>and</strong> Used Wholly) has decreased by 16%.<br />

The urban centers of the coastal strip has welded together in a metropolitan continuum that<br />

extends to approximately 15% of the region, from Caserta (North Campania) to Battipaglia<br />

(South Campania), but hosts 72% of the population of the Campania Region. The result was the<br />

abnormal consumption <strong>and</strong> waste of soil in the more fertile areas, as well as the loss of rural<br />

l<strong>and</strong>scapes of inestimable value. Several factors contributed to this unbalanced outcome: the<br />

inadequacy of public policies, a lack of participation of the public opinion, the influence of<br />

organized lawlessness (Totò 1952; di Gennaro, 2005).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Migliozzi et al. 2010. L<strong>and</strong>-use management <strong>and</strong> changes in Campania Region (Southern Italy)<br />

521<br />

Since 2004 a series of legislative measures have outlined the regional legal instruments capable<br />

of inverting this trend by offering an integrated strategy for the management <strong>and</strong> restoration of<br />

the urban, rural <strong>and</strong> forest ecosystems,. Within the framework of the l<strong>and</strong>ascape restoration the<br />

Piano Territorale Regionale (PTR) <strong>and</strong> the the Piano <strong>Forest</strong>ale Generale (PFG) was<br />

promulgated. Among the objectives of the latter are the protection, conservation <strong>and</strong><br />

enhancement of ecosystems, forest resources <strong>and</strong> hydro-geological aspects, <strong>and</strong><br />

theimprovement of the socio-economic conditions for the populations of the hill <strong>and</strong> mountains<br />

areas.<br />

The study of 10 state forests in the Campania Region was an applied research aimed to highlight<br />

both local <strong>and</strong> regional l<strong>and</strong>scape changes, for a more appropriate definition of the PFG<br />

recently approved by the Regional Council of Campania.<br />

In a first step the present work focuses on the change at l<strong>and</strong>scape scale occurring in the last 50<br />

years in Campania. Then, physiographic units were employed as a basis to describe the 10 state<br />

forests of the Campania Region. The main problems of sustainable forest management <strong>and</strong><br />

agro-forestry in Campania are highlighted in relation to actions proposed by the PRT <strong>and</strong> PFG.<br />

2. Methodology<br />

Two different approaches were followed: At a l<strong>and</strong>scape level, carriers responsible for the<br />

transformation of Campania’s l<strong>and</strong>scape agroforestry were identified, covering the last fifty<br />

years. A l<strong>and</strong>-systems map was used as background to focus on ten different forests,<br />

highlighting their status <strong>and</strong> management perspectives.<br />

2.1 L<strong>and</strong>scape changes towards 40 years<br />

The analysis of l<strong>and</strong>–use changes was carried out comparing, in GIS environment (ArcGIS -<br />

ESRI inc.), the l<strong>and</strong>-use map of ltaly (CNR TCI 1956-1960) with the Corine l<strong>and</strong> cover map<br />

(level IV – 2000 EU) integrated by the map of rural soil use, edited by the National Institute of<br />

Agricultural Economics (INEA CEC-2001), which allowed a more accurate cartographic<br />

restitution of farml<strong>and</strong>. Comparison was made with reference to physiographic areas (L<strong>and</strong>s<br />

systems) characterised by a particular combination of environmental factors (climate,<br />

morphology <strong>and</strong> soil) which can influence at the medium <strong>and</strong> long term sustainable use of<br />

resources by the man.<br />

Finally, the cartographic analysis data was integrated with socio-economic statistics <strong>and</strong> census<br />

(a more accurate description of the used approach is available at www.risorsa.info). The<br />

implementation of geographic database in GIS environment, has assumed the reference of all<br />

cartographic documents to a single geographic projection system (Gauss Boaga for Italy – Zone<br />

2).<br />

2.2 State <strong>Forest</strong>s characterisation<br />

A detailed survey of forest types was conducted within the 10 State <strong>Forest</strong>s of the Campania<br />

Region (Figure 1).<br />

State of art <strong>and</strong> bibliographic acquisition<br />

All the existing bibliographic <strong>and</strong> cartographicinformation on flora <strong>and</strong> vegetation <strong>and</strong> forest<br />

management plans related to the geographical areas where forests are located, was acquired <strong>and</strong><br />

archived according the guidelines for the sustainable management of forest <strong>and</strong> pastoral<br />

resources in Protected Areas. To assess the constraints on the management of the habitats of<br />

these areas, overlaps with the Site of Community Importance Special Protection Areas (SPAs)<br />

<strong>and</strong> Special Areas of Conservation (SACs), included in the network "Bioitaly-Nature 2000"<br />

were considered.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Migliozzi et al. 2010. L<strong>and</strong>-use management <strong>and</strong> changes in Campania Region (Southern Italy)<br />

522<br />

Figure 1: Location <strong>and</strong> list of the 10 State <strong>Forest</strong>s in Campania Region (Southern Italy)<br />

Ortho-Photo interpretation <strong>and</strong> mapping of the forest types<br />

The main forest formations were identified through the interpretation of digital orthophotos<br />

(1px / m resolution; max strain = 2m).It was possible to obtain a more precise classification of<br />

forest types in relation to the tone <strong>and</strong> texture of the pictures. The forest polygons were acquired<br />

directly from georeferenced digital orthophotos in a GIS environment (ArcView 3.2 ESRI inc.).<br />

Contextually, a geographic database for further processing was built. The polygon video scale<br />

capture was not less than 1:3,000 scale ratio.<br />

Field investigations<br />

In order to determine the main structural aspects of each forest type found on the digital<br />

orthophotos, a combined approach based on floro-vegetational, phyto-sanitary,<br />

phytosociological <strong>and</strong> physiognomical characteristics was carried out. Measurement on<br />

sampling areas, also documented by a detailed photographic survey, was performed by targeting<br />

plots, which allowed the compilation of field data sheets, useful for subsequent phases, <strong>and</strong> for<br />

outlining the current woodl<strong>and</strong> management .<br />

Field investigations regarded: 1) Biotic <strong>and</strong> abiotic environmental characteristics (st<strong>and</strong> site<br />

analysis); 2) Qualitative <strong>and</strong> quantitative features of the st<strong>and</strong>; 3) Productivity <strong>and</strong> nonproductivity<br />

functions of the woodl<strong>and</strong>s .<br />

3. Results<br />

L<strong>and</strong>scape changes through 40 years<br />

L<strong>and</strong> cover changes at regional scale can be summarized through a model based on three<br />

compartments (Figure 2):<br />

Figure 2: L<strong>and</strong> cover changes through 40 years in Campania Region. LAW (L<strong>and</strong> Used Wholly)<br />

contributes to almost all of the increase in urban areas<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Migliozzi et al. 2010. L<strong>and</strong>-use management <strong>and</strong> changes in Campania Region (Southern Italy)<br />

523<br />

The figure shows that LAW is affected by a net decrease of 175,322 ha, corresponding to 15.8%<br />

of the agricultural area used in 1960. This trend is opposed by the increase of 103,874 hectares<br />

of forest <strong>and</strong> shrub area (+43%), <strong>and</strong> of71,447 ha of urban areas (+321%).<br />

The overlay operation between l<strong>and</strong>-use dynamics maps <strong>and</strong> l<strong>and</strong>-systems map allowed to<br />

analyse the geographical distribution of l<strong>and</strong> cover dynamics in relation to: a) differences<br />

between the great systems of l<strong>and</strong> <strong>and</strong> the general physiographical partitions; b) differences<br />

within the major l<strong>and</strong> systems.<br />

Figure 3 shows a summary of these findings. 40% of LAW decrease is found on mountains,<br />

28% on hills, 10% in the volcanic complexes, 22% on plains.<br />

Figure 3: Result of the intersection between l<strong>and</strong>-systems <strong>and</strong> l<strong>and</strong>-use dynamics maps<br />

The net loss of LAW accounted 40% for urbanization <strong>and</strong> the remaining 60% for the<br />

recolonisation of natural vegetation by agricultural areas <strong>and</strong> grassl<strong>and</strong>s.<br />

<strong>Forest</strong>s State characterization<br />

The <strong>Forest</strong>s State of Campania are distributed in environments with different climate <strong>and</strong><br />

vegetation, between the the sea coast <strong>and</strong> the mountain belt (Figure). The territorial districts fall<br />

within both SCI <strong>and</strong> SPA areas of Natura 2000 network in the Regional Parks of Picentini,<br />

Partenio <strong>and</strong> Campi Flegrei <strong>and</strong> in the National Park of Cilento <strong>and</strong> Vallo di Diano.<br />

The forest type of the <strong>Forest</strong>s State consist in:<br />

1. coppices mostly exceeding the usual coppice cycle or on the way to conversion into<br />

high forest. They are classified into monospecific coppices of Quercus ilex, Q.<br />

pubescens Q. cerris <strong>and</strong>Castanea sativa coppices pure or mixed with oak, hornbeam,<br />

maples <strong>and</strong> ash trees;<br />

2. high st<strong>and</strong> forests consisting of deciduous trees (Fagus sylvatica, Quercus cerris <strong>and</strong><br />

other mesophilous species) <strong>and</strong> conifers such as Abies alba with groups of deciduous<br />

pioneers such as Betula pendula, Populus tremula <strong>and</strong> Alnus cordata. ) are introduced<br />

in different periods. Plantations of native woody species (Q cerris, Q. pubescens,<br />

Prunus avium) <strong>and</strong> alien species (e.g. Eucalyptus spp. Pseudotsuga menziesii.) were<br />

realised during the last 50-60 years ;<br />

3. coastal plain forests with extrazonal broadleaved species <strong>and</strong> riparian woodl<strong>and</strong>s<br />

dominated by willows <strong>and</strong> poplars.<br />

Due to the previous intensive forest management <strong>and</strong> grazing in the forest these st<strong>and</strong>s needs<br />

different sustainable management options. To each forest one or more functions (scientific,<br />

educational, touristic <strong>and</strong> recreational, protective <strong>and</strong> productive) can be assigned .<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Migliozzi et al. 2010. L<strong>and</strong>-use management <strong>and</strong> changes in Campania Region (Southern Italy)<br />

524<br />

The forest types identified refer to the following Categories <strong>and</strong> Types for Sustainable <strong>Forest</strong><br />

Management, Reporting <strong>and</strong> Policy (European forest types - EEA Technical Report No.<br />

9/2006):<br />

Table 1: Sustainable <strong>Forest</strong> Management types found in the Regional State <strong>Forest</strong>s<br />

7. Montane beech forest 7.3 Apennine‐Corsican montane beech forest<br />

8. Thermophilous deciduous forest<br />

9. Broadleaved evergreen forest<br />

10. Coniferous forests of the Mediterranean, Anatolian<br />

<strong>and</strong> Macaronesian regions<br />

12. Floodplain forest<br />

13. Non riverine alder, birch, or aspen forest<br />

14. Plantations <strong>and</strong> self sown exotic forest<br />

8.2 TURKEY OAK, HUNGARIAN OAK AND SESSILE OAK FOREST<br />

8.7 Chestnut forest<br />

8.1 Downy oak forest<br />

8.8 Other thermophilous deciduous forests<br />

9.1 Mediterranean evergreen oak forest<br />

9.5 Other sclerophlyllous forests<br />

10.6 Mediterranean <strong>and</strong> Anatolian fir forest<br />

12.1 Riparian forest<br />

12.2 Fluvial forest<br />

13.2 Italian alder forest<br />

13.4 Southern boreal birch forest<br />

13.5 Aspen forest<br />

14.1 Plantations of site‐native species<br />

14.2 Plantations of not‐site‐native species <strong>and</strong> self‐sown<br />

exotic forest<br />

4. Discussion<br />

A sustainable forest management of the 10 Regional State <strong>Forest</strong>s could be implemented also<br />

in the other forested areas. Among the factors of the degradation are the pressure of<br />

urbanization on environment <strong>and</strong> l<strong>and</strong>scape, pollution, widespread vulnerability of the soil<br />

(erosion <strong>and</strong> geological instability), coastal erosion <strong>and</strong> desertification, the increasing fire<br />

frequency, unsustainable touristic activities, lack of plans <strong>and</strong> forestry regulations at<br />

municipalitiy level, lack of forest grazing regulations, poor dissemination <strong>and</strong> applications of<br />

technological innovations in forestry. Nevertheless in these areas forested surfaces increased in<br />

the last years. They are located within protected areas (Natura 2000, Regional <strong>and</strong> National<br />

Parks, Marine Reserves <strong>and</strong>other Protected Areas) with high values of biodiversity (Figure 4) in<br />

relation to the variety of habitats.<br />

The main objectives to be pursued in the management of the State <strong>Forest</strong>s, contained in the PFG,<br />

concern the protection <strong>and</strong> improvement of biodiversity, in order to increase the stability <strong>and</strong><br />

bio-ecological features of the forest st<strong>and</strong>s <strong>and</strong> to preserve the soil from erosion <strong>and</strong><br />

degradation.<br />

The approval of the PTR <strong>and</strong> PFG provides the government of the Campania region the tools<br />

for a sustainable management of urban, rural <strong>and</strong> mountain areas according to guideline aimed<br />

at the protection of the l<strong>and</strong>scapes. Sustainably managed State <strong>Forest</strong>s could be used as pilot<br />

areas for monitoring natural forest processes, testing different management options <strong>and</strong><br />

disseminating the results. Their geographic distribution makes these areas representative of the<br />

major physiographic <strong>and</strong> ecological features present in the Campania Region.<br />

References<br />

di Gennaro A. (2005) (a cura di). Piani imperfetti. Il caso del piano urbanistico della provincia<br />

di Napoli. Clean Edizioni, Napoli.<br />

di Gennaro A., Innamorato F.P. 2005. La Gr<strong>and</strong>e Trasformazione – Il territorio rurale della<br />

Campania 1960-2000. Clean Edizioni, Napoli.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Migliozzi et al. 2010. L<strong>and</strong>-use management <strong>and</strong> changes in Campania Region (Southern Italy)<br />

525<br />

Mazzoleni S., Di Pasquale G., Mulligan M., Di Martino P., <strong>and</strong> Rego F. 2004 - Recent<br />

Dynamics of Mediterranean Vegetation L<strong>and</strong>scape. John Wiley & Sons, Ltd, 306 pp<br />

Mazzoleni S., Di Martino P., Di Pasquale G., Migliozzi A., Strumia S.(2001) "Urban forest<br />

management <strong>and</strong> slope stability: a study case in Napoli area". Atti della IUFRO<br />

Conference "Collecting <strong>and</strong> analyzing information for sustainable forest management <strong>and</strong><br />

biodiversity monitoring with special reference to mediterranean ecosystems" - Palermo 4-<br />

7 Dicembre 2001<br />

Motti R., Maisto A., Migliozzi A., Mazzoleni S. (2004).- Le trasformazioni del paesaggio<br />

agricolo e forestale dei Campi Flegrei nel XX secolo. Informatore Botanico Italiano 36<br />

(2) 577-583.<br />

Mingo A., Migliozzi A. & Maugeri G. 2008 - Integrated multi-scale approach to Mediterranean<br />

grassl<strong>and</strong>s. A case-study on the Nebrodi Mounts (Sicily). Options Mediterraneennes,<br />

Series A, No 79: 79-83. ISSN: 1016-121-X ISBN: 2-85352-378<br />

Acknowledgments<br />

The research project on the State <strong>Forest</strong>s of Campania Region has been supported by the<br />

Department of Agriculture of Campania Region <strong>and</strong> by the Consortium for Applied Research in<br />

Agriculture (CRAA). We thank dr.Gennaro Grassi, Dr. Daniela Lombardo <strong>and</strong> dr. Matilde<br />

Mazzaccara Assessorato all’Agricoltura REGIONE Campania - Settore <strong>Forest</strong>e Caccia e Pesca).<br />

We thank Francesco Paolo Innamorato (Risorsa s.r.l.) for the support to input data at regional<br />

scale <strong>and</strong> for graphic preparation.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L Porto et al. 2010. Naturalness <strong>and</strong> diversity of biotopes: their impact on l<strong>and</strong>scape quality 525 i<br />

Naturalness <strong>and</strong> diversity of biotopes: their impact on l<strong>and</strong>scape<br />

quality-a mathematical model<br />

Maria Luiza Porto 1* , Horst H. Wendel 2 , Jairo J. Zocche 3 , Gilberto G. Rodrigues 4 ,<br />

Marisa Azzolini 1 <strong>and</strong> Rogério Both 1 .<br />

1 Laboratory of L<strong>and</strong>scape Ecology, Department of Ecology, University Federal do Rio<br />

Gr<strong>and</strong>e do Sul, Av. Bento Gonçalves, 9500, PO Box: 15007, Porto Alegre, RS, Brazil<br />

2 Universität Ulm, D-89069 Ulm, Germany<br />

3 Laboratory for L<strong>and</strong>scape Ecology, Dept. of Biological Sciences, University of<br />

Extremo Sul Catarinense-UNESC, Av. Universitária, 1105, PO Box: 3167, Cricíuma,<br />

SC, Brazil<br />

4 Departament of Zoology, Center of Biological Sciences, University Federal de<br />

Pernambuco, Av. Prof. Morais Rego, 1235, Cidade Universitária. PO 50670-420,<br />

Recife, PE Brazil<br />

Abstract<br />

A novel mathematical tool serves to quickly grading l<strong>and</strong>scapes from the point of view of<br />

environ-mental value. Model is founded on five building blocks: (i) A l<strong>and</strong>scape is considered<br />

to be mosaic of biotic patches; (ii) each patch displays the characteristics of a single<br />

representative biotope out of a finite set of generic biotopes; (iii) generic biotope´s<br />

environmental value depends on its degree of na-turalness (flora, fauna) <strong>and</strong> diversity which are<br />

determined in field surveys; (iv) a l<strong>and</strong>scape´s environ-mental value is a simple function of the<br />

environmental values of its constituent biotopes; (v) the influence of shape <strong>and</strong> corrugated<br />

surface of a patch on the environmental value of the associated l<strong>and</strong>scape is reflected by an<br />

amplification factor. By accounting for the biotic properties (naturalness / biodiversity) <strong>and</strong> for<br />

the structure of the l<strong>and</strong>scape the model lends itself to comparative <strong>and</strong> trade off analyses in<br />

l<strong>and</strong> use management <strong>and</strong> urban development projects.<br />

Keywords: modelling, indices, naturalness of biotopes, diversity of biotopes, l<strong>and</strong>scape<br />

valuation<br />

1. Introduction<br />

According to Farina & Belgrano (2004), l<strong>and</strong>scape is the cognitive dimension of ecological<br />

complexity: Patterns <strong>and</strong> processes are distributed in a cognitive geographic space, the “ecofield”.<br />

The new paradigm allows to describe the biotic <strong>and</strong> abiotic characteristics of the<br />

physicoecological space. To underst<strong>and</strong> its foundations the notion of biotope is most helpful.<br />

Commonly conservational recommendations are based on a l<strong>and</strong>scape´s naturalness <strong>and</strong><br />

biodiversity. In view of what was said above this approach must be rolled out to its constituent<br />

biotopes.<br />

Efforts in describing l<strong>and</strong>scape´s naturalness <strong>and</strong> biodiversity were carefully reviewed in a<br />

recent paper of ours (Porto et al., 2010). Especially we pointed out that up to now theore-tical<br />

*Corresponding author: E-mail:mlporto@ecologia.ufrgs.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L Porto et al. 2010. Naturalness <strong>and</strong> diversity of biotopes: their impact on l<strong>and</strong>scape quality 525 ii<br />

models focused on structural aspects only (Papadimitriou, 2002, 2009). They did not link to the<br />

biota of the l<strong>and</strong>scape.<br />

Here we present a simple mathematical model which avoids the deficiencies stated above. It<br />

calculates one index only <strong>and</strong> thus lends itself to quick preselective l<strong>and</strong>scape assessment.<br />

Model uses empirical data <strong>and</strong> determines the ecological value of a l<strong>and</strong>scape through the<br />

degree of naturalness of its biotopes, i.e. structure <strong>and</strong> composition of their plant communities<br />

<strong>and</strong> structure <strong>and</strong> composition of the assemblage of their epigeic fauna. It accounts for<br />

indicators of disturbance, for the diversity of fauna <strong>and</strong> flora, <strong>and</strong> area <strong>and</strong> shape of the patches<br />

constituting the l<strong>and</strong>scape.<br />

2. Methodology<br />

In general a l<strong>and</strong>scape X is a cluster of patches distinct from the point of view of fauna <strong>and</strong><br />

flora. Mathematically the patches make up for a partition P of the l<strong>and</strong>scape X<br />

P = {X i , i = 1,2,…, n/Xi c X; Xi≠ Ø; X i ∩X j = Ø, i ≠ j; X 1 UX 2 U...UX n = X} , (1)<br />

where, ∩ is the intersection of two patches, U their union <strong>and</strong> X i c X indicates that X i is a subset<br />

of X. Each patch X i has an area A i <strong>and</strong> a perimeter P i . The sum of all subset areas is equal to the<br />

area of the l<strong>and</strong>scape,<br />

A tot = Σ A i . (2)<br />

In order to express the degree of naturalness of a biotope our model introduces an index of<br />

naturalness associated with that biotope. The part INT flor due to the flora prevailing in that<br />

biotope follows the equation<br />

INT flor = ((1 - IVI/IVI max ) + (1 - DR exot /DR maxexot )) . (3)<br />

IVI is the index of value of importance (Curtis <strong>and</strong> MacIntosh, 1951), for the species of the<br />

plant communities of the patch. DR exot is the relative density of exotic species of the biotope.<br />

In order to obtain the index of environmental value of the biotope as a function of its naturalness<br />

<strong>and</strong> its diversity, its index of naturalness is multiplied by the Shannon measure of diversity H',<br />

i.e. Peet (1974).<br />

IVA flor = INT flor * H´flor , (4)<br />

H´flor = -Σ (p i * lnp i ) flor . (5)<br />

(p i ) flor is the probability of occurrence of species i <strong>and</strong> ln denotes the natural logarithm. Eq. (4) is<br />

derived by pure arguments of mathematical plausibility: naturalness of a biotope whose species<br />

were planted is zero by definition <strong>and</strong> thus its environmental value equals zero. Nevertheless it<br />

still may exhibit a nonnegligible diversity, i.e. H´> 0. A good example for this is man-made<br />

soils where for purposes of restoration different species of herbs were planted. Equ. (4) accounts<br />

for this scenario.<br />

Design of the index of environmental value due to the biotope fauna (IVA faun ), follows a track<br />

similar to the one for deriving (IVA flor ): index of naturalness of the fauna of the biotope<br />

(INT faun ), thus is<br />

INT faun = ((1 - DR/DR max ) + (1- DR spider /DR maxspider )) (6)<br />

DR is the largest relative density of the epigeic fauna of a specific biotope <strong>and</strong> DR max its<br />

maximum value across all biotopes of the l<strong>and</strong>scape under consideration. DR spider is the relative<br />

density of spiders in a biotope. DR maxspider is the maximum relative spider density across all<br />

biotopes of the l<strong>and</strong>scape<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L Porto et al. 2010. Naturalness <strong>and</strong> diversity of biotopes: their impact on l<strong>and</strong>scape quality 525 iii<br />

As before multiplication of Eq. (6) by the diversity H' yields the index of environmental value<br />

of a biotope due to its fauna.<br />

IVA faun = INT faun *H´faun . (7)<br />

The index of environmental value of a circular biotope X i based on its naturalness <strong>and</strong> diversity,<br />

IVA oi , is the sum of its values IVA flor <strong>and</strong> IVA faun.<br />

IVA oi = IVA flor (i) + C*IVA faun (i). (8)<br />

C is the relative weight with which fauna <strong>and</strong> flora, respectively, contribute to the total<br />

environmental value of biotope i.<br />

To account for the fact that in general biotope Xi has a non-circular shape we introduce the<br />

factor<br />

(K F ) i = 2*(Π * A i ) ½ / P i (9)<br />

(K F ) i varies between 0 (biotope with extremely corrugated edge) <strong>and</strong> 1 (circular biotope). With<br />

this factor we define the index of environmental value of a biotope Xi to be<br />

IVA i = ((2 - (K F ) i ) * IVA oi (10)<br />

The index of environmental value of a l<strong>and</strong>scape based on the naturalness of its constituent<br />

biotopes, IVA tot , corresponds to the weight average of the environmental values of the biotopes:<br />

IVA tot = (A tot ) -1 * Σ A i * IVA i = (A tot ) -1 * Σ A i *((2 - (K F ) i ) * IVA oi (11)<br />

The weight with which biotope i enters this equation corresponds to its relative biotope area<br />

A i *(A tot ) -1 . (K F ) i is defined by Eq. (9), IVA oi through Eq. (8), <strong>and</strong> Σ represents a sum over all n<br />

biotopes that shape the l<strong>and</strong>scape.<br />

3. Result <strong>and</strong> Discussion<br />

We applied the model to a circular test sample of 200 m radius. Sample was chosen out of an<br />

area of survey located in the city of Treviso, State of Santa Catarina, southern Brazil. Its<br />

covering is hetero-genous: remnants of original vegetation (humid atlantic coastal forest), areas<br />

used for agriculture (livestock, fruit plantations), large waste deposits of coal mines <strong>and</strong> areas<br />

reforested with Eucalyptus. The area was found to be describable on the basis of five resilient<br />

natural biotopes: climax forest (GB1), early secondary forest (GB2), late secondary forest<br />

(GB3), eucalyptus plantation with forest regeneration (GB4), anthropogenic grassl<strong>and</strong>s (GB5),<br />

Fig 1. For each generic biotope we carried out phytossociological surveys <strong>and</strong> calculated the<br />

appropriate indices of importance, relative densities, <strong>and</strong> Shannon´s indices of diversity (Porto<br />

et al.. 2010) needed for executing Eq.(11). Structural information such as area <strong>and</strong><br />

circumference of the patches constituting the test sample was extracted from the digital image of<br />

the area of investigation. The resulting indices of environmental value, IVA, for the generic<br />

biotopes are shown in Fig. 2.<br />

According to our model, assumption of high diversity implying automatically high<br />

environmental quality is not valid (Fig. 2). On the scale of biotope for example, generic<br />

biotopes 4 <strong>and</strong> 5 exhibit relative high values of diversity both in fauna <strong>and</strong> flora. Their<br />

respective indices of environmental value though are lowest in the set of the five generic<br />

biotopes. This is because the corresponding indices of naturalness compensate the high values<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L Porto et al. 2010. Naturalness <strong>and</strong> diversity of biotopes: their impact on l<strong>and</strong>scape quality 525 iv<br />

of diversity, in our model. As a matter of fact our data revealed specific areas which originally<br />

were areas of native forest but now after reforestation represented Eucalyptus monocultures. In<br />

addition undergrowth had developed which showed a high degree of species diversity. For sure<br />

these areas show high diversity but are of low environmental value.<br />

References<br />

Curtis, J.T., McIntosh, R.P., 1951. An upl<strong>and</strong> forest continuum in the prairie-forest border<br />

region of Wisconsin. Ecology 32, 476-496.<br />

Farina, A. & Belgrano, A. (2004) The eco-field: A new paradigm for l<strong>and</strong>scapeecology.<br />

Ecological Research, 19, 107-110.<br />

Papadimitriou, F. (2002) Modelling indicators <strong>and</strong> indices of l<strong>and</strong>scape complexity: an<br />

approach using G.I.S. Ecological Indicators, 2, 17-25.<br />

Papadimitriou, F. (2009) Modelling spatial l<strong>and</strong>scape complexity using the Levenshtein<br />

algorithm. Ecological Informatics, 4, 48-55.<br />

Peet, R.K., 1974. The measurement of species diversity. Annu. Rev. Ecol. Syst. 5, 285-307.<br />

Maria Luiza Porto, Horst H. Wendel, Jairo J. Zocche, Gilberto G. Rodrigues, Marisa Azzolini<br />

<strong>and</strong> Rogério Both (2010) Index of Environmental Value: How Naturalness <strong>and</strong> Diversity<br />

of Biotopes Translate into L<strong>and</strong>scape Quality – A Mathematical Model, submitted to<br />

Journ. Ecol. Indic.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.L Porto et al. 2010. Naturalness <strong>and</strong> diversity of biotopes: their impact on l<strong>and</strong>scape quality 525 v<br />

Figure 1. Part of the l<strong>and</strong>scape for which empirical data were collected. The circular area represents the<br />

sample of diameter of 400 m used to test the mathematical model. Indicated are also the different types of<br />

soil covering of the 27 biotopes of the sample.<br />

7<br />

Numerical Value of H´, IVA<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

GB1 GB2 GB3 GB4 GB5<br />

Generic Biotope<br />

H´flor H´faun IVA<br />

Figure 2. Indices of environmental value (solid line) , IVA, Shannons´s diversity index of the fauna<br />

(dash-dotted line), H´faun , <strong>and</strong> the flora (dashed line), H´flor , respectively, for the five generic biotopes of<br />

the sample of Fig.1.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Rehor & V. Ondracek 2010. Application of modern, non traditional restoration methods on brown coal localities<br />

526<br />

Application of modern, non traditional restoration methods on brown<br />

coal localities of Czech Republic<br />

Michal Rehor 1* & Vratislav Ondracek 2<br />

1 Brown Coal Research Institute, Most, Czech Republic<br />

2 North Bohemian Mines, Chomutov, Czech Republic<br />

Abstract<br />

Over 70% brown coal reserves have been exploited in the North-Bohemian Basin today.<br />

Opencast mining of brown coal naturally led to vast l<strong>and</strong>scape damages. Therefore reclamation<br />

work has acquired a great significance. The research methodology of the areas of interests <strong>and</strong><br />

the reclamation works themselves described in this article arises from North Bohemian Mines<br />

locality reclamation philosophy. Apart from the earlier published methodology of fertilizable<br />

soils application used today in operation, experiments with filling areas left for natural<br />

succession <strong>and</strong> pilot application of power plant stabilizer <strong>and</strong> ash in phyto-toxic areas. The<br />

results are stated in the paper.<br />

Key words: Restoration, dump, geology, methodology<br />

1. Introduction<br />

The North-Western Bohemia region takes a unique position in the history of mining in<br />

the Czech Republic. Several ore districts are situated here, the welfare of which influenced the<br />

development of the Czech state in some periods. Some precious stones mining was also<br />

significant. Today the North-Bohemian Basin area is known for its largest Czech brown coal<br />

deposit. Whereas ore mining in the region extinguished <strong>and</strong> the other raw materials mining has<br />

relatively small significance, coal mining <strong>and</strong> the restoration of damages caused by mining<br />

activity still have an essential meaning.<br />

Severoceské doly, a.s. Chomutov (North Bohemian Mines, j.s.c. Chomutov), the<br />

greatest mining company of the Czech Republic, was founded within the privatization of the<br />

North-Bohemian Brown-Coal Mines concern privatization on 1.1.1994. Today the allotments of<br />

the share holding company include geological reserves of about 1200 million tonnes <strong>and</strong> the<br />

recoverable reserves amounting 700 million tonnes. The annual production amounts to ca 22<br />

million tonnes of brown coal representing almost 50% Czech market with this coal, <strong>and</strong> the<br />

annual volume of overburden stripping is 100 million m 3 of clayey rocks. This is currently the<br />

largest mining company in the Czech Republic. The reclamation of vast areas affected with<br />

brown coal mining is, of course, an important part of its activities.<br />

Current task of Severoceské doly, a.s. reclamations is to make reclamation works more<br />

efficient using locally available fertilizable rocks <strong>and</strong> other materials, <strong>and</strong> to be maximaly<br />

environmentally friendly. The core of the article is the characteristics of new reclamation<br />

methods used mainly for reclamation of sterile <strong>and</strong> phyto-toxical areas of both localities of<br />

Severoceské doly, a.s. <strong>and</strong> for the protection, research <strong>and</strong> relevant development of remarkable<br />

ecosystems arising in the dumps.<br />

* Corresponding author. Tel.: +420603935808<br />

Email address: rehor@vuhu.cz<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Rehor & V. Ondracek 2010. Application of modern, non traditional restoration methods on brown coal localities<br />

527<br />

2 . New Concept of Technical Reclamation in the North Bohemian Brown<br />

Coal Basin<br />

Mining <strong>and</strong> reclamation works of Severoceské doly, a.s. proceed in two considerably<br />

different geological areas. It concerns Nástup Tusimice Mines with the mining locality of the<br />

Libous mine <strong>and</strong> the Bílina Mines with the mining locality of the Bílina mine. Based on the<br />

different type of overburden rocks in both mining localities different requirements for the areas<br />

reclamation appear. In case of the Bílina Mines the main issue is the occurrence of extremely<br />

acid phyto-toxical areas (high contents of coal – ca 5%), in the case of the Nástup Tusimice<br />

Mines the main issue is the occurrence of sterile areas (high content of physical clay). The<br />

research methodology of the areas of interests <strong>and</strong> the reclamation works themselves described<br />

in this article arises from Severoceské doly, a.s. locality reclamation philosophy. It is based on<br />

the knowledge of overburden minerals properties <strong>and</strong> detailed survey of each reclaimed sites<br />

provided in co-operation with the Severoceské doly, a.s., Výzkumný ústav pro hnedé uhlí, a.s.<br />

(Research Institute of Brown Coal, j.s.c. Most), Výzkumný ústav meliorací a ochrany půdy<br />

Praha (Research Institute of Ameliorations <strong>and</strong> Soil Protection Praha) <strong>and</strong> Zemedelská<br />

universita Praha (Agricultural University of Prague). Experiments have been started recently<br />

with areas left for natural succession, with application of power plant stabiliser (this product is<br />

described in greater detail in chapter five) <strong>and</strong> power plant ash in phyto-toxic areas apart from<br />

the methodology of fertilizable soils application earlier published by (Ondrácek 2003),<br />

(Safárová 2003) <strong>and</strong> being used in operation today.<br />

3. Application of Fertilizable Rocks in Reclaimed Localities of the North Bohemian<br />

Brown Coal Basin<br />

Top soil, loess <strong>and</strong> loess loam, marlite <strong>and</strong> bentonite are the most important rocks used<br />

for reclaimed purposes in the localities of Severoceské doly, a.s.<br />

The application of fertilizable rocks is the most efficient in areas consisting of highly<br />

arenaceous to phyto-toxical rocks. In this case marlaceous minerals or bentonite minerals are<br />

applied in the amount 3000-3500 m 3 .ha -1 with the following homogenation (inmixture) or cross<br />

ploughing from 0,5 to 0,6 m. Surface overlay in the area of interest with loess loam <strong>and</strong> the<br />

thickness to 0,5 m is an option of this procedure. The application of organic substances<br />

(composts) with the adjusted ratio C : N (25) in the amount 400 t.ha -1 , embedded into 0,30-0,50<br />

m reclaimed surface of the dump <strong>and</strong> the follow-up two-year preparatory agricultural cycle<br />

(growing plants for green manure) is requested as an additional measure.<br />

The stated methodology was successfully used in a lot of sites of the Bílina Mine.<br />

Anthropogenic soil profile created by bentonite application in the Strimice dump, an<br />

anthropogenous soil profile created with the application of marl in the Radovesice dump <strong>and</strong> the<br />

anthropogenous soil profile created with the application of loess loams at the inner dump of the<br />

Bílina Mine can be used as an example. Several results are shown in the table No. 1.<br />

4. Filling Areas Left for Natural Succession at Radovesice Dump<br />

The areas left for natural succession have been tentatively filled in the areas where<br />

functional ecosystems spontaneously started to develop under specific conditions where<br />

protection <strong>and</strong> research of some biological, geological <strong>and</strong> paleontological phenomenon is<br />

necessary <strong>and</strong> where future access to public may be assumed within the overall concept of the<br />

locality reclamation. The selection of these areas proceeds based on the research of dumps.<br />

After the area is filled <strong>and</strong> plotted into the planning maps detail research is provided based on<br />

which entry documentation is created. Thus long-term research of the area starts evaluating its<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Rehor & V. Ondracek 2010. Application of modern, non traditional restoration methods on brown coal localities<br />

528<br />

pedological <strong>and</strong> biological development. The article briefly characterizes the areas filled at the<br />

Radovesice dump which is the largest reclaimed dump of the Bílina Mines <strong>and</strong> also the largest<br />

dump in the Czech Republic.<br />

Based on the extended survey of the non-cultivated part of the dump (ca 670 ha)<br />

consisting of terrain mapping, evaluation of top soil profile sampled with a boring bar <strong>and</strong><br />

laboratory analyses of the selected samples two quite large areas were selected to be left for<br />

natural succession in the year 2004.<br />

Succession area 1 (32 ha) was selected in the southern part of the dump. Heterogeneous<br />

dump mixture of brown clay, grey claystone <strong>and</strong> grey s<strong>and</strong> claystone with higher content of<br />

brown clay was a prevailing mineral type. Brown-grey kaolinitic- illitic clays appear. In the<br />

eastern part of the area s<strong>and</strong>y minerals are more significantly represented creating natural border<br />

of the area. A lot of natural water bodies <strong>and</strong> wetl<strong>and</strong> occur here (see figure No. 1).<br />

Succession area 2 (20) ha was selected in the northern part of the dump. The mineral<br />

consumption of the upper horizon is similar as in case of the area No 1. Southern border of the<br />

area is created with “s<strong>and</strong> dunes”. Two large natural water bodies <strong>and</strong> several small water<br />

bodies <strong>and</strong> wetl<strong>and</strong>s occur here. Some small water bodies verge in the wetl<strong>and</strong> during the year.<br />

Plant <strong>and</strong> animal species composition was evaluated in both areas. It is recommended to<br />

leave the area for natural development without reclamation impacts. The area should be<br />

monitored in the future <strong>and</strong> serve for research purposes. It is interesting to monitor the way how<br />

several kinds (mainly from plant l<strong>and</strong>) comply to a certain, not very typical environment for<br />

these species. With respect to the area situation it will also serve as a natural corridor for the<br />

animals movement in necessary technical works in the surrounding part of the dump.<br />

Pedological properties of the prevailing mineral type are shown in the table No. 2.<br />

5. Experimental Application of Power Plant Stabilizer in Inner Dump of Bílina<br />

Mine<br />

In the year 2006 the application of power plant ash application to various types of dump<br />

soils was tested in the inner dump of the Bílina mine. Extremely acid (sterile from the<br />

reclamation point of view) coal claystones from the Ledvice preparation plant which created<br />

large phytotoxic area was one of these types. Stabilizer from the Ledvice power plant was used<br />

for the experiment (the desulphurization product here). The objective of the work was to<br />

evaluate the chances to make the technical reclamation of phyto-toxic area more efficient – see<br />

(Rehor 2004).<br />

The used stabilizer is the product of desulphurisation in Ledvice plant. It contains<br />

CaSO 3 (cca 50%), CaCO 3 (cca 25%), Ca(OH) 2 (cca 20%), <strong>and</strong> others less significant<br />

components. Dangerous contents of risky trace elements were not found out. It is very alcalic,<br />

that is why the possibility was tested of its application on extremely acidic phyto-toxic area of<br />

coal claystones.<br />

The values of the soil reaction in the water extract were, for the above stated coal<br />

claystones, usually about 3,8 – 4,5. After the application of power plant stabilizer in the amount<br />

600 t.ha -1 soil reaction of the resulted mixture was about 9-10. From whence it followed the<br />

need to optimize the stabilizer doses to reach the soil reaction of the resulted mixture ca 6,5 –<br />

7,5.<br />

In the task solution first samples of clean power plant stabilizer <strong>and</strong> coal claystone in<br />

which the value of soil reaction in soil extract were taken. Then an experimental area with the<br />

dimensions 0,5 x 0,5 m was established to which various doses of stabilizers were embedded<br />

<strong>and</strong> then the soil reaction of the mixture detected. The embedded dose of stabilizer was<br />

calculated per 1 hectare of the area. The results are stated in the table No. 3 below. The samples<br />

were taken immediately after the application of stabilizer <strong>and</strong> after every 2 years.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Rehor & V. Ondracek 2010. Application of modern, non traditional restoration methods on brown coal localities<br />

529<br />

Optimum dosing ranges between 100 – 300 t.ha -1 , for further application the dose 200<br />

t.ha -1 can be recommended. The mixtures A-F were tested for the presence of risk trace<br />

elements, all samples comply with the valid legislation of the Czech Republic. The experiment<br />

proved substantial improvement of the sterile coal claystones properties <strong>and</strong> the method is, with<br />

respect to great production of stabilizer, very prospective. Before its practical use other tests in<br />

larger areas will be necessary.<br />

6. Experimental Application of Power Plant Ash in Brezno Dump<br />

Application of power plant ash from the Tusimice power plant was tested in the<br />

experimental areas of the Brezno dump. It concerns the external dump of the Libous. mine.<br />

Experimental areas were found on the yellow overburden clays from the Libous mine. Some<br />

other details are stated by (Rehor 2004).<br />

These rocks are homogeneous, strongly viscid <strong>and</strong> cast. With respect to extremely<br />

unfavourable physical properties <strong>and</strong> water regime they are not suitable for reclamation use.<br />

Their hydrophysical soil properties do not change even after longer period of depositing on the<br />

surface of the dump. They create permanently strained soil structure with unfavourable<br />

infiltration capabilities.<br />

The target of the experiment was to improve grain composition of the upper horizon of<br />

the tested areas. Power plant ash in the amount 200 t.ha -1 was put on three areas with the<br />

acreage 1 ha <strong>and</strong> embedded with cultivator to the top horizon. Then the grain composition of the<br />

former yellow clays <strong>and</strong> the generated mixture was detected. The results of the grain analyses<br />

are stated in the table No. 4. Apart from that the sample with the applied ash was tested for the<br />

presence of risky trace elements. It has been confirmed that it complies with the valid legislation<br />

of the Czech Republic.<br />

The results show significant improvement of the grain compositions of the mineral after<br />

the application of power plant ash. While former yellow clay can be ranked, from the grain size<br />

point of view, to the clay category, the mixture with power plant ash can be ranked as a clayey<br />

soil. But equal embedding of ash into clay is an issue. Gross description of samples however<br />

showed that ash creates rather single isolated clumps.<br />

Despite of these issues the method is prospective, further research will focus on<br />

preparing optimum methodology of embedding power plant ash into the upper horizon of the<br />

reclaimed localities.<br />

7. Discussion<br />

The shareholding company Severoceské doly, a.s. Chomutov is today the largest mining<br />

company in the Czech Republic. Big volumes of stripped rocks <strong>and</strong> the extent of areas<br />

designated for reclamation require the application of new, efficient reclamation methods.<br />

The application of fertilizable minerals remains the basic methods of technical<br />

reclamation proving its success at the localities of Strimice, Radovesice <strong>and</strong> the inner dump of<br />

the Bílina mine. But the first attained results of the research prove that some new progressive<br />

methods can become a significant complement of this method. Application of power plant<br />

stabilizer into extremely acid phyto-toxic minerals in Bílina Mines could be a prospective<br />

method where some non-traditional erosion control measures applied well. In the Nástup<br />

Tusimice Mines areas of interest, in case of resolving the issue of homogenation, the application<br />

of power plant ash into the heavy grain clays can be significant. Thanks to great morphological<br />

<strong>and</strong> geological variety of non-reclaimed areas of the Severoceské doly, a.s. there is enough<br />

space for filling the areas left for natural succession the target of which is the protection of<br />

unique ecosystems developing in the dumps.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Rehor & V. Ondracek 2010. Application of modern, non traditional restoration methods on brown coal localities<br />

530<br />

The conclusions stated in this article are documented with the results of samples taken<br />

in Strimice, Radovesice, inner dump Bílina, Brezno <strong>and</strong> the overburden cuts of the Bílina Mine.<br />

This work has been prepared with the support of the Czech Ministry of Education,<br />

Youth <strong>and</strong> Physical Training within the research work No. MSM 4456918101 “Research of<br />

Physical <strong>and</strong> Chemical Properties of Substances Affected with Coal Mining <strong>and</strong> Use <strong>and</strong> Their<br />

Impact on Environment in North-Western Bohemia Region“.<br />

References<br />

Ondrácek, V., Rehor, M., Safárová, M., Lang, T.: Historie, Gegenwart und Perspektiven der<br />

rekultivierung auf dem Gebiet des Bergbaubetriebes Doly Bílina Surface Mining<br />

magazine - Braunkohle: p. 90-100, ISSN 0931 – 3990, No 1, 2003, Deutschl<strong>and</strong><br />

Ondrácek, V., Cermák, P., Rehor, M.: Reducing Danger of Wind <strong>and</strong> Water<br />

Erosion <strong>and</strong> Their Application in Bílina Mines Localities Coal, ores, geological survey<br />

magazine: p.12-16, ISSN 1210 – 7697, No 5, Prague 2006<br />

Rehor, M., Safárová, M., Ondrácek: Application of Some Coal Treatment Products for<br />

Reclamation of Localities in the North Bohemian Basin 21th. Pittsburg Coal Conference,<br />

Osaka, 2004<br />

Safárová, M., Rehor, M., Lang, T.: Application of Modern Restoration Methods in Vicinity of<br />

Bilina Mines Journal of the Polish Mineral Engineering Society, 2: p. 9-27, PL ISSN<br />

1640 – 4920, No 2, 2003<br />

Table No. 1: Features of Reclaimed Soil Profile after Application of Fertilizable<br />

Rocks<br />

N C ox CaCO 3 pH/ Acceptable nutrients Adsorbing capability<br />

H (mg.kg -1 2 0<br />

) mmol/100 g (%)<br />

Taking<br />

interval<br />

(m)<br />

(%) (%) (%)<br />

P K Mg S T V<br />

Strimice locality<br />

0,00-,60 0,07 1,24 0,98 6,79 10 190 102 13,4 19,7 68<br />

0,60-0,90 0,07 0,68 9,93 8,23 1 218 949 36,3 36,3 100<br />

0,90-1,10 0,05 2,94 0,24 4,50 1 103 304 3,1 8,2 39<br />

Radovesice locality<br />

0,00-0,30 0,12 0,83 9,41 7,93 1 148 96 16,8 16,8 100<br />

0,30-0,90 0,09 0,44 39,42 8,25 0 112 26 15,6 15,6 100<br />

0,90-1,10 0,10 1,03 1,068 7,90 0 196 269 7,9 7,9 100<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Rehor & V. Ondracek 2010. Application of modern, non traditional restoration methods on brown coal localities<br />

531<br />

Table No. 2: Pedological Properties of Upper Horizon of Area Left f or Natural<br />

Succession<br />

Taking N Cox. CaCO 3 pH/ Acceptable nutrients Adsorbing capability<br />

interval<br />

H 2 O (mg.kg -1 ) mmol/100 g (%)<br />

P K Mg S T V<br />

(m) (%) (%) (%)<br />

0,00-0,90 - 5,6 0,1 2,5 2 11 27 0 12 0<br />

Table No. 3: Recommended Doses of Stabilizer <strong>and</strong> Final Values of Soil Reaction<br />

Sample Stabilizer dose / 1 ha<br />

(t)<br />

pH/H 2 O<br />

- after application<br />

pH/H 2 O<br />

- after 2 years<br />

Clean<br />

- 12,1<br />

stabilizer<br />

A 600 9,63 8,52<br />

B 500 8,86 8,00<br />

C 400 8,57 7,80<br />

D 300 8,12 7,50<br />

E 200 7,25 7,20<br />

F 100 7,00 7,00<br />

Coaly clay 0 4,11<br />

Figure 1:<br />

Demonstration of Small Water Bodies <strong>and</strong> Wetl<strong>and</strong> Spontaneously<br />

Developed in Successive Areas of Radovesice<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Russell et al. 2010. Developing models <strong>and</strong> processes to aid decision support for integrated l<strong>and</strong> management<br />

532<br />

Developing models <strong>and</strong> processes to aid decision support for integrated<br />

l<strong>and</strong> management in the Canadian boreal forest<br />

Jonathan Russell 1 , Ellie Prepas 2,3* , Gordon Putz 4 , Daniel W. Smith 3 & James Germida 4<br />

1<br />

Millar Western <strong>Forest</strong> Products Ltd., Whitecourt, Alberta, Canada<br />

2<br />

Lakehead University, Thunder Bay, Ontario, Canada<br />

3<br />

University of Alberta, Edmonton, Alberta, Canada<br />

4<br />

University of Saskatchewan, Saskatoon, Saskatchewan, Canada<br />

Abstract<br />

<strong>Forest</strong>ed l<strong>and</strong>s encompass nearly half of the surface area of Canada <strong>and</strong> are exposed to dem<strong>and</strong>s<br />

from a growing human population. Industrial forested l<strong>and</strong>s support multiple uses (e.g., forestry,<br />

oil <strong>and</strong> gas <strong>and</strong> urban development) within the same space <strong>and</strong> time. Planning processes need to<br />

consider these cumulative, long-term impacts. A novel forest management planning initiative is<br />

presented that considers forested l<strong>and</strong>scape conditions within the context of climate change <strong>and</strong><br />

cumulative impacts from forestry related <strong>and</strong> non-forestry related disturbances, as well as<br />

habitat supply <strong>and</strong> water quantity <strong>and</strong> quality. These measures are analyzed under various<br />

models to produce an optimum long-term forest strategy that satisfies current economic drivers<br />

<strong>and</strong> the new paradigm of intrinsic values that are naturally contained within the forest.<br />

Keywords: integrated l<strong>and</strong> management, cumulative effects assessment<br />

1. Introduction<br />

Canada is approximately 10 million km 2 in area, of which >30% is boreal forest (Fig. 1). More<br />

than 90% of the boreal forest is classified as Provincial Crown L<strong>and</strong> (under the jurisdiction of<br />

Provincial Governments in the name of the Crown) (NRC 2009). Within provinces, mechanisms<br />

exist to develop ~20-year leases on Crown L<strong>and</strong> to allow forest products companies to establish,<br />

grow <strong>and</strong> harvest timber. <strong>Forest</strong> Management (FM) plans must be developed by the companies<br />

<strong>and</strong> approved by the Province (usually every 10 years), with a 200-year forest harvest volume<br />

projection. There is usually a specific spatial harvest sequence for the first 5-10 years, which<br />

details where forestry operations will be undertaken. In Alberta (Fig. 1), the FM planning<br />

process, managed through the Ministry of Sustainable Resource Development, is centered on<br />

timber supply analysis <strong>and</strong> has a limited focus on issues like water, wildlife <strong>and</strong> the physical<br />

environment. Further, it virtually ignores the issues of Integrated L<strong>and</strong> Management (ILM) <strong>and</strong><br />

Cumulative Effects Assessment (CEA), as well as oil <strong>and</strong> gas activities, which can have as big a<br />

footprint as the forest industry. <strong>Forest</strong> operations on adjacent FM areas are also not considered,<br />

eliminating any potential to mitigate large l<strong>and</strong>scape issues that encompass multiple FM areas.<br />

This study is based primarily in the FM area (productive l<strong>and</strong> base 2950 km 2 ) leased by Millar<br />

Western <strong>Forest</strong> Products Ltd., a small pulp <strong>and</strong> solid wood products company. The research<br />

group embarked upon a process that would help push the intent of FM planning by engaging the<br />

company, government <strong>and</strong> researchers in an ILM approach (Russell et al. 2008). This FM plan<br />

would be based on science, data <strong>and</strong> state-of-the-art modelling <strong>and</strong> would address issues like<br />

* Corresponding author. Tel.: 807-343-8623 - Fax: 807-343-8116<br />

Email address: eprepas@lakeheadu.ca<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Russell et al. 2010. Developing models <strong>and</strong> processes to aid decision support for integrated l<strong>and</strong> management<br />

533<br />

ILM <strong>and</strong> CEA. The plan would demonstrate that it was feasible to produce an ILM/CEA<br />

structure, models could be developed or adapted, <strong>and</strong> outcomes could be tested over time.<br />

2. Methodology<br />

In 1995, Millar Western developed permanent vegetation sample plots that focused on tree<br />

growth, as well as free-to-maneuver flying space, bat crevices, tree lichen cover <strong>and</strong> downed<br />

woody debris. Additional research included creel censuses, thinning trials to evaluate tree<br />

growth <strong>and</strong> habitat use, tree retention in clearcut st<strong>and</strong>s, animal surveys <strong>and</strong> habitat supply<br />

model verification. In 1997, expert panels were assembled to develop models for: l<strong>and</strong>scape<br />

projection under various climate change scenarios; coarse <strong>and</strong> fine filter habitat supply <strong>and</strong>;<br />

surface water quantity <strong>and</strong> quality. Research outcomes <strong>and</strong> the mechanisms developed would be<br />

incorporated into a timber supply model as constraints against the allowable cut.<br />

2.1 L<strong>and</strong>scape Projection<br />

The L<strong>and</strong>scape group used the Special Report on Emissions Scenario A1 (IPCC 2001) to predict<br />

impacts on the forest of climate change, namely a doubling in atmospheric CO 2 concentrations,<br />

<strong>and</strong> an increase of 7ºC in temperature <strong>and</strong> 8% in precipitation between 1999 <strong>and</strong> 2100 (Millar<br />

Western <strong>Forest</strong> Products 2007). To project climate conditions, the CCSR-NIES <strong>Global</strong><br />

Circulation Model (Ozawa et al. 2001) was selected. To downscale climate data, local weather<br />

was adjusted with global circulation model data. The FORECAST model (Kimmins et al. 1999)<br />

was used to predict forest growth <strong>and</strong> the Athabascan Plains L<strong>and</strong>scape model was used to<br />

predict forest succession changes (Yamasaki et al. 2008). Nine scenarios were tested <strong>and</strong><br />

biodiversity <strong>and</strong> productivity were evaluated within these l<strong>and</strong>scape projections (Yamasaki et al.<br />

2008). The study area sits at the northern limit of the Aspen Parkl<strong>and</strong> ecoregion <strong>and</strong> research<br />

has suggested that climate change may result in the conversion of forested l<strong>and</strong> to aspen<br />

parkl<strong>and</strong> (Hogg <strong>and</strong> Hurdle 1995). Therefore this analysis included prediction of how much<br />

forested l<strong>and</strong> would shift to aspen parkl<strong>and</strong> under climate change Scenario A1 (IPCC 2001).<br />

Daily fire weather indices were calculated once weather streams under historical <strong>and</strong> climate<br />

change conditions were obtained. The Fire Behavior Prediction System (<strong>Forest</strong>ry Canada Fire<br />

Danger Group 1992) was used to derive daily values for the Fine Fuel Moisture Code, Build Up<br />

Index <strong>and</strong> Fire Weather Index for the period 1 January 1961 to 31 December 1990 <strong>and</strong> the<br />

corresponding 30-year period under climate change. This fire record, which corresponds to the<br />

30-year baseline most commonly used in climate research, serves as the basis for fire modelling<br />

with the Athabascan Plains L<strong>and</strong>scape Model. To simulate 200 years of fires, the model cycles<br />

through this 30-year record almost seven cycles. Fire duration, extent, frequency <strong>and</strong> seasonality<br />

were all statistically modelled as dependents of the Fire Behavior Prediction indices above <strong>and</strong><br />

annual, seasonal <strong>and</strong> daily measures of temperature, precipitation <strong>and</strong> wind. A link detected<br />

among fire occurrence, weather <strong>and</strong> human population was also included in the analysis.<br />

As an example of a dominant non-forestry related human disturbance in the FM area, patterns of<br />

oil <strong>and</strong> gas development were analyzed <strong>and</strong> future developments were simulated. The number of<br />

current wells was determined from the forest inventory. It was also possible to generate a time<br />

series of number of new well sites annually, <strong>and</strong> the area each represents. This was also<br />

completed for pipelines <strong>and</strong> seismic lines that are projected to being developed in the FM area.<br />

2.2 Habitat Supply<br />

Coarse filter analysis was divided into two areas within the Biodiversity Assessment Project<br />

(BAP): ecosystem diversity <strong>and</strong> l<strong>and</strong>scape configuration (Van Damme et al. 2003). Ecosystem<br />

diversity analysis assumes that silvicultural practices will modify the distribution of ecosystems<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Russell et al. 2010. Developing models <strong>and</strong> processes to aid decision support for integrated l<strong>and</strong> management<br />

534<br />

across space <strong>and</strong> time. To monitor changes in the composition of the forecasts, BAP tracks the<br />

proportion of habitat types <strong>and</strong> diversity of the forest using the following metrics. 1) Areaweighted<br />

age is a single value at each time step during the simulation, indicating the average<br />

age of the entire forest. 2) Tree species distribution separates the forest into hardwood st<strong>and</strong>s,<br />

hardwood-dominated mixedwood st<strong>and</strong>s, softwood st<strong>and</strong>s <strong>and</strong> softwood-dominated mixedwood<br />

st<strong>and</strong>s. It provides an indication of the proportion of the FM area expected to support each<br />

habitat type at each time step. 3) Species presence indicates the extent of coverage of each tree<br />

species over the l<strong>and</strong>scape. 4) Species dominance takes into account both tree species presence<br />

<strong>and</strong> relative dominance. 5) Habitat diversity was computed using a matrix showing similarity<br />

between habitat types as a weighting factor. The habitat diversity index considers the relative<br />

position of broad habitat types using a rating of similarity between the habitat types as a<br />

weighting factor. The diversity equation generates a single unitless value between 0 <strong>and</strong> 1 at<br />

each time step; 0 represents a very uniform l<strong>and</strong>scape <strong>and</strong> 1 indicates the most diverse<br />

l<strong>and</strong>scape possible. Incorporated into the index are both the number of habitat types present<br />

within the FM area <strong>and</strong> the proportion of the l<strong>and</strong>scape covered by each habitat type.<br />

<strong>L<strong>and</strong>scapes</strong> containing many habitat types distributed evenly across the area are considered<br />

more diverse than those dominated by one habitat type, yet containing small portions of others.<br />

Within the l<strong>and</strong>scape configuration analysis, habitat types were used as class attributes because<br />

they can be weighted by contrast. As well, different levels of distinction among habitats can be<br />

used following the classification hierarchy. The l<strong>and</strong>scape configuration analysis was comprised<br />

of several types of biostatistical analyses:<br />

• Patch – areas of similar characteristics based on age <strong>and</strong> species composition.<br />

• Edge – mean edge contrast index <strong>and</strong> contrast weighted edge length.<br />

• Core area – The impact of edge on wildlife was expected to be linearly related to the<br />

abruptness of the habitat structure change at the edge <strong>and</strong> therefore the buffer width would<br />

change with contrast between adjacent habitat patches.<br />

• Adjacency – It was expected that the spatial distribution of habitat types would differ in a<br />

managed scenario relative to a natural scenario. Consequently the proportion of adjacencies<br />

might be different. Many species use a combination of different habitats to fulfill their<br />

needs; therefore the adjacency of these habitats is important.<br />

• Nearest neighbor – It was understood that a population using an isolated habitat is highly<br />

prone to local extinction. In addition, an organism using dispersed habitat patches may not<br />

be able to defend a sufficiently large territory from which to extract its needed resources.<br />

The nearest neighbor metric gives an indication of the dispersion of similar habitat types.<br />

Fine filter analysis was undertaken by developing species-specific habitat supply models. It was<br />

recognized that the species selected would comprise an imperfect representation of the large<br />

array of species that occupy the FM area. The selection process was based on the following<br />

premises. 1) It is not possible to create models for all species. 2) Coarse filter analysis can<br />

account for habitat requirements for many species. 3) Models created for a carefully selected list<br />

of species will adequately represent the habitat needs of many other species. 4) Terrestrial<br />

vertebrates will be selected because: they use a large range of forest features; are good<br />

indicators of change; are the subjects of concern by the public <strong>and</strong>; approaches for analyzing<br />

forests in terms of vertebrate habitat potential are relatively well developed. Species were<br />

selected that represented the following classes: large terrestrial carnivorous mammals, large<br />

ungulates, medium-sized herbivores/omnivorous mammals, medium-sized carnivorous<br />

mammals, small mammals, raptors, birds in the order Gallinaceae <strong>and</strong> passerines (perching<br />

birds). The following criteria were used, with the numbers relating to the weighting scheme for<br />

each element: sensitivity to disturbance – 4; species status – 3; ability to monitor the species – 3;<br />

habitat specificity – 2; special habitat elements – 2; functionally essential species – 2; l<strong>and</strong>scape<br />

configuration – 2; socioeconomic value – 2 <strong>and</strong>; available information – 1.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Russell et al. 2010. Developing models <strong>and</strong> processes to aid decision support for integrated l<strong>and</strong> management<br />

535<br />

Habitat supply models were developed for each of 17 species selected to evaluate the potential<br />

of the FM area to provide suitable habitat (Van Damme et al. 2003). The models define habitat<br />

suitability based on the provision of habitat elements required for survival <strong>and</strong> reproduction.<br />

Within the models, special habitat element models were developed to characterize changes in<br />

condition (i.e. abundance, density) of habitat elements through forest succession <strong>and</strong> disturbance.<br />

Specific (i.e. canopy closure, tree height), general (i.e. perches, hiding cover) <strong>and</strong> habitat uses<br />

(i.e. food) were included in the analysis. Habitat supply models projected suitability based on<br />

home range size for each species <strong>and</strong> were used to assess each forest harvest scenario at 5-year<br />

time steps. Additionally, the special habitat elements were analyzed within the habitat supply<br />

model to determine a subset of elements that were most critical for the 17 species. These were<br />

then associated with silvicultural practices, seral stages <strong>and</strong> dominant species groups <strong>and</strong><br />

embedded within the timber supply model as constraints.<br />

2.3 Water<br />

The <strong>Forest</strong> Watershed <strong>and</strong> Riparian Disturbance (FORWARD) project developed hydrologic<br />

models at two spatial scales (first- <strong>and</strong> third-order watershed) to predict how forest harvesting<br />

would change precipitation-normalized stream runoff at the watershed outlet (i.e. corrected for<br />

both watershed area <strong>and</strong> precipitation inputs). This variable, termed a runoff coefficient, was<br />

predicted for a given set of watershed attributes under forested conditions using a variant of the<br />

Soil <strong>and</strong> Water Assessment Tool (SWAT) (Arnold et al. 1998). The data to populate these<br />

runoff coefficient models were collected before <strong>and</strong> after experimental harvest from treatment<br />

<strong>and</strong> reference watersheds. Data included streamflow at a high temporal resolution, as well as<br />

improved digital elevation model, slope, soil <strong>and</strong> wetl<strong>and</strong> coverage, <strong>and</strong> watershed boundaries.<br />

SWAT simulations are still underway to compare measured to predicted runoff coefficients with<br />

time after harvest <strong>and</strong> to calibrate the model for future FM planning cycles (Watson et al. 2008).<br />

Water quality is being incorporated into SWAT modelling, as well as artificial neural network<br />

modelling (Nour et al. 2006). The FORWARD project established indicators <strong>and</strong> ranges of<br />

acceptability in changes in streamflow based upon changes in runoff coefficients (Prepas et al.<br />

2008). Thus, potential changes to streamflow could be given equal or varying weight as<br />

compared to other forest values during the development of a FM plan.<br />

3. Results <strong>and</strong> Discussion<br />

The structure of the FM planning analysis is hierarchical. The first task was to set the future<br />

l<strong>and</strong>base condition that included climate <strong>and</strong> vegetation change, human population change,<br />

wildfire response to these two conditions <strong>and</strong> oil <strong>and</strong> gas activities (Fig. 2). This produced a<br />

l<strong>and</strong>scape change outside the normal projection (i.e. these issues are not included in a st<strong>and</strong>ard<br />

timber supply model). These conditions act as forest l<strong>and</strong>base constraints: over the st<strong>and</strong>ard<br />

planning period of approximately 200 years these conditions enact significant changes in the<br />

forest l<strong>and</strong> base <strong>and</strong> forest growth patterns. This then produced future l<strong>and</strong> base series on which<br />

we could overlay a timber supply analysis. This analysis includes the constraints of biodiversity<br />

(BAP) <strong>and</strong> water (FORWARD) (Fig 2). This was conducted for multiple future scenarios at two<br />

basic levels of projection: multiple scenarios of each component part (e.g., different oil <strong>and</strong> gas<br />

scenarios could be incorporated into the model leaving all other components static) <strong>and</strong><br />

projections with various elements turned on or off, dependent upon what multiple effects needed<br />

to be focused upon for analysis. In this manner any number of iterations could be undertaken to<br />

arrive at an underst<strong>and</strong>ing of how each element affects the system at certain levels or how<br />

various combinations of elements can affect the system based on projected future scenarios.<br />

The system described above is a logical evolution from the st<strong>and</strong>ard timber supply analysis<br />

employed in much of Canada to a more holistic model concept that looks at all issues facing a<br />

forest state. It is critical to model elements of varying forest values outside of traditional<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Russell et al. 2010. Developing models <strong>and</strong> processes to aid decision support for integrated l<strong>and</strong> management<br />

536<br />

economic considerations when developing a timber supply analysis, because these cannot alone<br />

be addressed by best management practices that are based on good intentions, with little data,<br />

science or modelling expertise behind them. It was found that these values can be incorporated<br />

into the process as post-projection analyses of the scenarios. However, the optimal approach is<br />

to include the elements as constraint conditions within a planning system that weighs <strong>and</strong><br />

analyzes all elements at the same time. This approach demonstrates that the technical capacity is<br />

readily available to produce a true ILM/CEA plan. With all the pressures facing the forest l<strong>and</strong><br />

state, there is a definite need to produce plans such as these <strong>and</strong> what is needed now is the<br />

political <strong>and</strong> industrial will to carry such endeavors forward.<br />

4. Acknowledgments<br />

The FORWARD project (2001 to 2006) was funded by the Natural Sciences <strong>and</strong> Engineering<br />

Research Council of Canada, Millar Western <strong>Forest</strong> Products Ltd., the Canada Foundation for<br />

Innovation, Blue Ridge Lumber Inc., Alberta Newsprint Company, V<strong>and</strong>erwell Contractors<br />

(1971) Ltd., the Living Legacy Research Program, <strong>and</strong> the Ontario Innovation Trust. We thank<br />

Janice Burke (Lakehead University) for review of earlier drafts of this manuscript.<br />

5. References<br />

Arnold, J.G., Srinivasan, R., Muttiah, R.S. <strong>and</strong> Williams, J.R., 1998. Large area<br />

hydrologic modeling <strong>and</strong> assessment. Part I: Model development. Journal of the<br />

American Water Resources Association, 34: 73-89.<br />

<strong>Forest</strong>ry Canada Fire Danger Group, 1992. Development <strong>and</strong> structure of the Canadian<br />

<strong>Forest</strong> Fire Behavior Prediction System. Ottawa, Ontario: <strong>Forest</strong>ry Canada. 64 p.<br />

Hogg, E. <strong>and</strong> Hurdle, P., 1995. The aspen parkl<strong>and</strong> in western Canada: A dry climate<br />

analogue for the future boreal forest Water, Air <strong>and</strong> Soil Pollution, 82: 391-400.<br />

IPCC (Intergovernmental Panel on Climate <strong>Change</strong>), 2001. Climate change 2001:<br />

Impacts, adaptation <strong>and</strong> vulnerability. J.J. McCarthy, O.F. Canziani, N.A. Leary, D.J.<br />

Dokken <strong>and</strong> K.S. White (Eds.). Cambridge, UK: Cambridge University Press.<br />

Kimmins, J.P., Mailly, D. <strong>and</strong> Seely, B., 1999. Modelling forest ecosystem net primary<br />

production: the hybrid simulation approach used in FORECAST. Ecological<br />

Modelling, 122: 195-224.<br />

Millar Western <strong>Forest</strong> Products, 2007. Millar Western <strong>Forest</strong> Products Ltd. Detailed<br />

<strong>Forest</strong> Management Plan 2007–2016. Edmonton, Alberta: The <strong>Forest</strong>ry Corp. 451 p.<br />

NRC (Natural Resources Canada), 2009. The state of Canada’s forests. Ottawa,<br />

Ontario: Natural Resources Canada. 64 p.<br />

Nour, M.H., Smith, D.W., Gamal El-Din, M. <strong>and</strong> Prepas, E.E., 2006. Neural networks<br />

modelling of streamflow, phosphorus, <strong>and</strong> suspended solids: application to the<br />

Canadian Boreal forest. Water Science <strong>and</strong> Technology, 53(10): 91-99.<br />

Ozawa, T.N., Mori, S.E., Umaguti, A.N., Sushima, Y.T., Akemura, T.T., Akajima, T.N.,<br />

Beouchi, A.A. <strong>and</strong> Imoto, M.K., 2001. Projections of future climate change in the<br />

21 st century simulated by the CCSR/NIES CGCM under the IPCC SRES scenarios.<br />

In: T. Matsuno <strong>and</strong> H. Kida (Eds.). Present <strong>and</strong> future of modeling global<br />

environmental change: Toward integrated modeling. Tokyo, Japan: Terrapub: 15-28.<br />

Prepas, E.E., Putz, G., Smith, D.W., Burke, J.M. <strong>and</strong> MacDonald, J.D., 2008. The<br />

FORWARD Project: Objectives, framework <strong>and</strong> initial integration into a Detailed<br />

<strong>Forest</strong> Management Plan in Alberta. <strong>Forest</strong>ry Chronicle, 84: 330-337.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J. Russell et al. 2010. Developing models <strong>and</strong> processes to aid decision support for integrated l<strong>and</strong> management<br />

537<br />

Russell, J.S., Smith, D.W., Putz, G. <strong>and</strong> Prepas, E.E., 2008. Science <strong>and</strong> the industrial<br />

planning process in the western Canadian boreal forest: a case study. Journal of<br />

Environmental Engineering <strong>and</strong> Science, 7 (Suppl. 1): 1-12.<br />

Van Damme, L., Russell, J.S., Doyon, F., Duinker, P.N., Gooding, T., Hirsch, K.,<br />

Rothwell, R. <strong>and</strong> Rudy, A., 2003. The development <strong>and</strong> application of a decision<br />

support system for sustainable forest management on the Boreal Plain. Journal of<br />

Environmental Engineering <strong>and</strong> Science, 2 (Suppl. 1): 23-34.<br />

Watson, B.M., McKeown, R.A., Putz, G. <strong>and</strong> MacDonald, J.D., 2008. Modification of<br />

SWAT for modelling streamflow from forested watersheds on the Canadian Boreal<br />

Plain. Journal of Environmental Engineering <strong>and</strong> Science, 7 (Suppl. 1): 145-159.<br />

Yamasaki, S., Duchesneau, R., Doyon, F., Russell, J. <strong>and</strong> Gooding, T., 2008. Making<br />

the case for cumulative impacts assessment: Modelling the potential impacts of<br />

climate change, harvesting, oil <strong>and</strong> gas, <strong>and</strong> fire. <strong>Forest</strong>ry Chronicle, 84: 349-368.<br />

Figure 1: Map of Canada showing the approximate location of boreal forest <strong>and</strong> the location of the study<br />

area in the province of Alberta (inset).<br />

Figure 2: Schematic of the hierarchical <strong>Forest</strong> Management planning analysis. PFMS: Preferred<br />

<strong>Forest</strong> Management Strategy.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Section 7<br />

Management <strong>and</strong> sustainability of changing l<strong>and</strong>scapes


P. Angelstam & M. Elbakidze. 2010. Sustainable forest management, multi-stakeholder governance <strong>and</strong> spatial planning<br />

539<br />

Sustainable forest management requires multi-stakeholder governance<br />

<strong>and</strong> spatial planning: Kovdozersky state forest management unit in<br />

northwest Russia<br />

Per Angelstam * <strong>and</strong> Marine Elbakidze<br />

Swedish University of Agricultural Sciences, Faculty of <strong>Forest</strong> Sciences,<br />

School for <strong>Forest</strong> Engineers, PO Box 43, SE-739 21 Skinnskatteberg, Sweden<br />

Abstract<br />

Large-scale clear-felling of naturally dynamic forests during the Soviet period has left many<br />

remote forest l<strong>and</strong>scapes in the Russian Federation with limited wood resources for decades to<br />

come. Ideas about rural development based on forest non-wood goods, ecosystem services as<br />

well as natural <strong>and</strong> cultural l<strong>and</strong>scape values are thus emerging. To underst<strong>and</strong> stakeholders’<br />

needs for regional development based on both use <strong>and</strong> non-use values of forest l<strong>and</strong>scapes we<br />

interviewed 31 stakeholders from private, public <strong>and</strong> civil sectors in the Kovdozersky state<br />

forest management unit in southernmost Murmansk oblast. While about half of the stakeholders<br />

were confined to the Kovdozersky forest management unit (~500,000 ha), the spatial scales of<br />

stakeholder activities ranged from local villages to the entire catchment of Kovda River in the<br />

Russian Federation <strong>and</strong> Finl<strong>and</strong> (~2,600,000 ha). To avoid future negative externalities <strong>and</strong><br />

risks for conflicts there is a need to (1) communicate the state <strong>and</strong> trends about sustainability<br />

dimensions of forest l<strong>and</strong>scapes at multiple levels, (2) encourage collaborative learning<br />

processes about natural resource management based on principles of adaptive governance <strong>and</strong><br />

adaptive management, <strong>and</strong> (3) plan at multiple spatial scales to satisfy <strong>and</strong> reconcile the needs<br />

<strong>and</strong> interests of different l<strong>and</strong>scape stakeholders. The development of traditional <strong>and</strong> new<br />

products from forest l<strong>and</strong>scape’s goods, ecosystem services <strong>and</strong> values, <strong>and</strong> of local <strong>and</strong><br />

regional forest governance requires exchange of experiences with development initiatives<br />

toward an integrated l<strong>and</strong>scape approach in other regions <strong>and</strong> countries.<br />

Keywords: spatial planning, natural resource governance, rural development, boreal forest<br />

1. Introduction<br />

The circumboreal boreal forest ecoregion is an important provider of natural resources in terms<br />

of wood, minerals <strong>and</strong> hydroelectric energy supporting human welfare <strong>and</strong> quality of life. Being<br />

relatively little impacted by anthropogenic change, there is opportunity for biodiversity<br />

conservation including viable populations, ecological integrity <strong>and</strong> resilience (Angelstam et al.<br />

2004). Northern forest ecosystem also form a biomass sink (Myneni et al. 2001).<br />

According to international <strong>and</strong> national policy documents sustainable forest management (SFM)<br />

aims at satisfying economic, ecological <strong>and</strong> socio-cultural values, <strong>and</strong> should be based on the<br />

principles of sustainable development as good governance including representation of actors<br />

<strong>and</strong> stakeholders (Lammerts van Buren <strong>and</strong> Blom 1997; Mayers <strong>and</strong> Bass 2004). The effects of<br />

multi-level external factors from local <strong>and</strong> regional to national <strong>and</strong> global levels that affect<br />

ecosystems, including climate change, <strong>and</strong> social systems, need to be understood (Angelstam et<br />

al. 2005). Finally, the actors, stakeholders <strong>and</strong> organisations exercising government <strong>and</strong><br />

governance need to be well informed to connect forests <strong>and</strong> markets.<br />

* Corresponding author. Tel.: +46 (0)70-2444971 e-mail: per.angelstam@smsk.slu.se<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Angelstam & M. Elbakidze. 2010. Sustainable forest management, multi-stakeholder governance <strong>and</strong> spatial planning<br />

540<br />

A framework for building capacity for multi-level governance <strong>and</strong> planning towards sustainable<br />

forest l<strong>and</strong>scapes is to view l<strong>and</strong>scapes as interconnected social-ecological systems. This is<br />

often termed “l<strong>and</strong>scape approach”, meaning that that there is a need to exp<strong>and</strong> the area for<br />

planning from st<strong>and</strong>s <strong>and</strong> local areas, <strong>and</strong> to support participation of representative stakeholders<br />

(World <strong>Forest</strong>ry Congress 2009). The social dimensions involve the institutions <strong>and</strong> all<br />

stakeholders involved with the use of natural resources. Thus, sustainable l<strong>and</strong>scapes are<br />

integrated systems encompassing diverse cultural, natural <strong>and</strong> social functions through balanced<br />

governance empowering the involvement of all actors <strong>and</strong> stakeholders (Norton 2005; Baker<br />

2006). To describe the ecosystem providing renewable natural resources, its composition,<br />

structure <strong>and</strong> function at multiple scales need to be understood. Similarly, the actions of the<br />

social system’s actors <strong>and</strong> stakeholders from different sectors <strong>and</strong> governance levels should be<br />

mapped. Angelstam et al. (2003) coined the term two-dimensional gap analysis to better<br />

underst<strong>and</strong> policy implementation processes in integrated social-ecological systems.<br />

Many special initiatives have appeared globally that aim at implementing sustainability <strong>and</strong><br />

sustainable development on the ground (Axelsson et al. 2008). The Model <strong>Forest</strong> (MF) concept,<br />

which originated in Canada in the beginning of the 1990s, is one example (IMFN 2008). A MF<br />

can be seen as a social process aimed at forming a partnership <strong>and</strong> a platform for discussing <strong>and</strong><br />

solving a wide spectrum of issues related to sustainable management in a forest l<strong>and</strong>scape by<br />

implementing new ideas approved by MF partners, <strong>and</strong> thus developing the adaptive capacity in<br />

an area to deal with uncertainty <strong>and</strong> change (LaPierre 2002).<br />

This study focuses on the southern part of the Murmansk region in the northwest of the Russian<br />

Federation. Mining <strong>and</strong> forest logging began in the late 19th century <strong>and</strong> hydro-electrical power<br />

stations were built from the 1950s. While mining <strong>and</strong> hydropower production continues,<br />

forestry has declined dramatically (Elbakidze et al. 2007). As a consequence, rural settlements<br />

are declining, <strong>and</strong> people rely on local use of a wide range of wood <strong>and</strong> non-wood resources,<br />

<strong>and</strong> emerging tourism based on natural <strong>and</strong> cultural values (e.g., Lehtinen 2006). To adapt<br />

governance <strong>and</strong> management of forest l<strong>and</strong>scapes to this situation a partnership of stakeholders<br />

was created in 2006 in the Kovdozersky forest management unit in the southernmost part of<br />

Murmansk region termed the Kovdozersky Model <strong>Forest</strong> (MF) (Elbakidze et al. 2007). We<br />

mapped l<strong>and</strong>scape stakeholders <strong>and</strong> their use of l<strong>and</strong>scape goods, services <strong>and</strong> values in the MF<br />

area. Our study shows the need for spatial planning at four spatial scales to communicate the<br />

present, past <strong>and</strong> future states <strong>and</strong> trends of forest goods, ecosystem services <strong>and</strong> values to<br />

different stakeholders at multiple level of organisation. We discuss the need to promote new<br />

forms of forest l<strong>and</strong>scape governance promote collaboration among stakeholders.<br />

2. Methodology<br />

2.1. Study area<br />

The Kovdozersky state forest management unit (400,626 ha) is located in the southern part of<br />

the Murmansk region in the north-west of the Russian Federation (Elbakidze et al. 2007).<br />

Geographically, it occupies the lower part of the Kovda river catchment, which has its<br />

headwaters on both sides of the Russian-Finnish border, <strong>and</strong> flows to the K<strong>and</strong>alaksha Bay of<br />

the White Sea, <strong>and</strong> borders with the Karelian Republic in the south. <strong>Forest</strong>s are dominated by<br />

Scots pine (Pinus sylvestris) at lower altitude <strong>and</strong> Norway spruce (Picea abies) on higher hills.<br />

There are eight settlements <strong>and</strong> many ab<strong>and</strong>oned logging villages from the period of<br />

exploitative logging in the area. Local people in rural settings have retained their traditional<br />

l<strong>and</strong>-use practices, which are based on forest resources. In 2006 the total number of people<br />

living in the area was about 15,000, <strong>and</strong> the population density was 3.7 people sq. km<br />

(Elbakidze et al. 2007).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Angelstam & M. Elbakidze. 2010. Sustainable forest management, multi-stakeholder governance <strong>and</strong> spatial planning<br />

541<br />

2.2. Methods<br />

To identify the stakeholders of the Kovdozersky MF <strong>and</strong> their use of forest l<strong>and</strong>scapes, both<br />

qualitative <strong>and</strong> quantitative methods were used 2006-2008. A total of 36 open-ended qualitative<br />

interviews were conducted with stakeholders in the Kovdozersky MF. All interviews were taken<br />

face to face <strong>and</strong> lasted between 60 <strong>and</strong> 180 minutes. The interviews focused on the natural<br />

resource use profiles of stakeholders from civil, private <strong>and</strong> public sector actors, trends in forest<br />

management <strong>and</strong> transport logistics, as well as economic <strong>and</strong> social development in the area of<br />

the Kovdozersky MF. To quantify the state <strong>and</strong> trend of economic, ecological <strong>and</strong> socio-cultural<br />

dimensions of SFM in the Kovdozersky MF the interviews were complemented with analyses of<br />

socio-economic statistical data from the All-Russian Population Census (2002), <strong>and</strong> archives of<br />

local <strong>and</strong> regional administrations. To make a survey of the products derived from different<br />

kinds of natural resources we divided them into use values <strong>and</strong> non-use values (Merlo <strong>and</strong><br />

Croitoru 2005). Direct use values include (1) consumptive (e.g., wood <strong>and</strong> non-wood goods) as<br />

well as (2) non-consumptive direct use values in terms of l<strong>and</strong>scape quality for recreation <strong>and</strong><br />

tourism. Indirect use values include ecosystem services such as watershed protection, water<br />

purification <strong>and</strong> carbon sequestration. Non-use values are closely linked to conservation<br />

interests of the l<strong>and</strong>scape.<br />

3. Results<br />

3.1. L<strong>and</strong>scape stakeholders<br />

In total we identified 31 stakeholders operating in the area of the Kovdozersky management unit.<br />

There were 13 l<strong>and</strong> leasers, 9 forest contractors, 6 other forest users, <strong>and</strong> 3 potential forest<br />

contractors whose rights to use forests was under negotiation with the state as a main l<strong>and</strong> <strong>and</strong><br />

forest owner (Figure 1). Stakeholders from all societal sectors used l<strong>and</strong>scape goods, services<br />

<strong>and</strong> values to create products through different kind of activities. However, private sector<br />

stakeholders were the main group (55% of the total number of stakeholders in the area). The<br />

representatives of the private sector were mainly small-scale forest logging companies, tourist<br />

enterprises, <strong>and</strong> an agricultural company. The smallest group was civil sector stakeholders,<br />

represented by the local gardening society. Stakeholders from the public sector managed the<br />

most of the l<strong>and</strong>, forests <strong>and</strong> water in the area of Kovdozersky MF. This group included the<br />

Kovdozersky state forest management unit, the hydropower stations, <strong>and</strong> the protected areas.<br />

The distribution of stakeholders among different groups indicated that it was an interest from<br />

the private sector to develop business. According to the interviews with the stakeholders, we<br />

concluded that the business interests were very diverse, ranging from small-scale forest industry<br />

based on local forest resources, tourism (nature-based, sport, fishing <strong>and</strong> hunting), maintenance<br />

<strong>and</strong> restoration of fish population, small scale farming <strong>and</strong> fisheries, <strong>and</strong> construction of bioenergy<br />

production facilities.<br />

3.2. Spatial planning scales<br />

The spatial levels of stakeholder activities covered a wide spectrum, from international to local.<br />

However, almost 50% of all stakeholders focused their activity in the Murmansk region, i.e. the<br />

regional level. Based on interviews we found that four spatial scales of stakeholders’ activities<br />

need to be considered in different types of spatial planning (Table 1). This means that although<br />

the different stakeholders which leased or used forest l<strong>and</strong> in the same forest management unit,<br />

the creation of products from l<strong>and</strong>scapes’ goods, services <strong>and</strong> values took place at different<br />

spatial scales.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Angelstam & M. Elbakidze. 2010. Sustainable forest management, multi-stakeholder governance <strong>and</strong> spatial planning<br />

542<br />

4. Discussion<br />

4.1. Spatial planning at multiple scales<br />

Today the main planning unit in the Kovdozersky MF is the 400,000 ha state forest<br />

management unit. An important task is to support co-existence of forest l<strong>and</strong> leasers operating at<br />

different spatial scales with planning for multiple uses within <strong>and</strong> among leasing areas. A total<br />

of 17 stakeholders lease parts of this management unit for wood harvesting, hunting <strong>and</strong><br />

recreation businesses. To secure sustainable use of forest l<strong>and</strong>scape goods, ecosystem services<br />

<strong>and</strong> values their spatial distribution needs to be mapped. Assessment is needed of wood<br />

resources for timber <strong>and</strong> bioenergy, habitat suitability for game <strong>and</strong> fish, nature <strong>and</strong> culture<br />

tourism, hydro-electric development <strong>and</strong> for nature conservation. Some users have also a need<br />

for regional trans-boundary planning at the scale of the entire Kovda river catchment (2,610,000<br />

ha) in Murmansk oblast, Republic of Karelia <strong>and</strong> Finl<strong>and</strong>. Three examples are hydro-electric<br />

development, conservation of the last intact forest areas in Fennosc<strong>and</strong>ia, as well as tourism<br />

linked to the Finnish <strong>and</strong> Russian cultural heritage. Given the importance of hydroelectric<br />

development for l<strong>and</strong>scape stakeholders <strong>and</strong> fish populations a catchment perspective is logic. It<br />

would be appropriate to include the entire Kovda river catchment into the Kovdozersky MF.<br />

4.2. Need for multi-stakeholder partnership for sustainability<br />

Implementing policies about sustainability on the ground is highly dependent on biophysical<br />

conditions, the environmental history, <strong>and</strong> the systems of government <strong>and</strong> governance.<br />

Stakeholders <strong>and</strong> actors at multiple levels use goods, services <strong>and</strong> values of forest l<strong>and</strong>scapes to<br />

develop products at several temporal <strong>and</strong> spatial scales. Consequently, when developing local<br />

<strong>and</strong> regional governance arrangements towards SFM policy implementation on the ground,<br />

stakeholders have to collaborate <strong>and</strong> develop capacity for learning to deal with uncertainties <strong>and</strong><br />

risks in governance <strong>and</strong> management of forest l<strong>and</strong>scapes. The development of collective action<br />

towards sustainable forest l<strong>and</strong>scapes differs in different situations <strong>and</strong> places. Different<br />

stakeholders may also have different interests <strong>and</strong> needs for taking part in collective action<br />

(Elbakidze et al. 2010). In the Russian Federation the first MF initiative appeared in 1994, <strong>and</strong><br />

in 2006 there were already five MFs. According to the “Initiative Network of Russian Model<br />

<strong>Forest</strong>s”, the Russian MFs are long-term projects, which develop on the basis of generally<br />

recognized international <strong>and</strong> Russian principles of SFM (Elbakidze <strong>and</strong> Angelstam 2008). At<br />

the end of 2007, the Russian Federation’s <strong>Forest</strong> Agency, inspired by the MF concept, planned<br />

to create 31 Model <strong>Forest</strong>s in addition to the five existing ones (Zheldak 2008). The vision was<br />

that the suite of MFs should represent all forest zones in the Russian Federation, <strong>and</strong> would<br />

become examples of SFM based on Russian <strong>and</strong> international experiences (Elbakidze <strong>and</strong><br />

Angelstam 2008). However, since then no official information has been provided about any<br />

governmental actions towards development of a MF network in Russia. Nevertheless, given the<br />

interest to base regional development on forest resources in Russia, applying integrated<br />

l<strong>and</strong>scape approaches, such as the MF concept, to support SFM implementation remains is an<br />

urgent task (World <strong>Forest</strong>ry Congress 2009).<br />

4.3. Knowledge production using transdisciplinary approaches<br />

To realise the vision of SFM a societal learning process it is crucial to explore different<br />

l<strong>and</strong>scape approaches in order to develop (i) an accounting system as a “map <strong>and</strong> a compass”<br />

that tells natural resource managers, policy-makers, media, authorities exercising governance,<br />

<strong>and</strong> the general public where we are going, <strong>and</strong> (ii) ways of establishing societal arenas for local<br />

<strong>and</strong> regional governance as a “gyroscope” that allows us to make informed decisions based on<br />

knowledge. This requires new forms of knowledge production that integrate different<br />

disciplines as well as academic <strong>and</strong> non-academic actors, i.e. transdisciplinary approaches (e.g.,<br />

Gibbons et al. 1994).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Angelstam & M. Elbakidze. 2010. Sustainable forest management, multi-stakeholder governance <strong>and</strong> spatial planning<br />

543<br />

References<br />

Angelstam, P., Boutin, S., Schmiegelow, F., Villard, M.-A., Drapeau, P., Host, G., Innes, J.,<br />

Isachenko, G., Kuuluvainen, M., Mönkkönen, M., Niemelä, J., Niemi, G., Roberge, J.-M.,<br />

Spence, J., Stone, D. 2004. Targets for boreal forest biodiversity conservation – a rationale<br />

for macroecological research <strong>and</strong> adaptive management. Ecological Bulletins, 51: 487-509.<br />

Angelstam, P., Mikusinski, G., Rönnbäck, B.-I., Östman, A., Lazdinis, M., Roberge, J.-M.,<br />

Arnberg, W., Olsson, J. 2003. Two-dimensional gap analysis: a tool for efficient<br />

conservation planning <strong>and</strong> biodiversity policy implementation. Ambio, 33(8): 527-534.<br />

Angelstam, P., Kopylova, E., Korn, H., Lazdinis, M., Sayer, J.A., Teplyakov, V. <strong>and</strong> Törnblom,<br />

J. 2005. Changing forest values in Europe. In: Sayer, J.A., Maginnis, S. (Eds) <strong>Forest</strong>s in<br />

l<strong>and</strong>scapes. Ecosystem approaches to sustainability. Earthscan: 59-74.<br />

Axelsson, R., Angelstam, P., Elbakidze, M. 2008. L<strong>and</strong>scape approaches to sustainability. In:<br />

Frostell, B., Danielsson, Å., Hagberg, L., Linnér, B.-O. & Lisberg Jensen E. (Eds.) Science<br />

for sustainable development. The social challenge with emphasis on the conditions for<br />

change. VHU, Uppsala, pp. 169-177.<br />

Baker, S. 2006. Sustainable development. Routledge, London <strong>and</strong> New York. 245 pp.<br />

Elbakidze, M., Angelstam, P. <strong>and</strong> Axelsson, R. 2007. Sustainable forest management as an<br />

approach to regional development in the Russian Federation: state <strong>and</strong> trends in<br />

Kovdozersky Model <strong>Forest</strong> in the Barents region. Sc<strong>and</strong>inavian Journal of <strong>Forest</strong> Research,<br />

22: 568-581.<br />

Elbakidze, M. <strong>and</strong> Angelstam, P. 2008. Model <strong>Forest</strong>s in the Russian Federation’s northwest: a<br />

view from the outside. Sustainable <strong>Forest</strong> Management, 17(1): 39-47. (In Russian).<br />

Elbakidze,M., Angelstam, P., S<strong>and</strong>ström, C. <strong>and</strong>. Axelsson, R. 2010. Multi-stakeholder<br />

collaboration in Russian <strong>and</strong> Swedish Model <strong>Forest</strong> initiatives: adaptive governance<br />

towards sustainable forest l<strong>and</strong>scapes Ecology <strong>and</strong> Society in press.<br />

Gibbons, M., L. Limoges, H. Nowotny, S. Schwartman, P. Scott, <strong>and</strong> Trow, M. 1994. The New<br />

Production of Knowledge. The Dynamics of Science <strong>and</strong> Research in Contemporary<br />

Societies. Sage, London.<br />

IMFN. 2008. Model <strong>Forest</strong> development guide. International Model <strong>Forest</strong> Network Secretariat.<br />

Lammerts van Buren, E.M., Blom, E.M. 1997. Hierarchical framework for the formulation of<br />

sustainable forest management st<strong>and</strong>ards. Principles, criteria, indicators. Tropenbos<br />

Foundation, Backhuys Publishers, AH Leiden, The Netherl<strong>and</strong>s. 82 pp.<br />

LaPierre, L. 2002. Canada’s Model <strong>Forest</strong> program. The <strong>Forest</strong>ry Chronicle, 78, 613-617.<br />

Lehtinen, A. A. 2006. Postcolonialism, multitude, <strong>and</strong> the politics of nature. On the changing<br />

geographies of the European North. University Press of America, Lanham.<br />

Mayers, J., Bass, S. 2004. Policy that works for forests <strong>and</strong> people. Real prospects for<br />

governance <strong>and</strong> livelihoods. Earthscan, London, UK. 324 pp.<br />

Merlo, M., Croitoru, L. 2005. Concepts <strong>and</strong> methodology: a first attempt towards quantification.<br />

In: Merlo, M., Croitoru, L. (Eds.) Valuing Mediterranean forests. Towards total economic<br />

value. CABI Publishing: 17-36.<br />

Myneni, R. B., Tucker, C. J., Kaufmann, R. K., Kauppi, P. E., Liski, J., Zhou, L., Alexeyev, V.,<br />

Hughes, M. K. 2001. A large carbon sink in the woody biomass of Northern forests. PNAS,<br />

98(26):14784–14789.<br />

Norton, B.G. 2005. Sustainability. A philosophy of adaptive ecosystem management. The<br />

University of Chicago press, Chicago <strong>and</strong> London. 607 pp.<br />

World <strong>Forest</strong>ry Congress. 2009. <strong>Forest</strong> Development: A Vital Balance, Findings <strong>and</strong> Strategic<br />

Actions. Findings <strong>and</strong> Strategic Actions, see:<br />

http://foris.fao.org/meetings/download/_2009/xiii_th_world_forestry_congress/misc_docum<br />

ents/wfc_declaration.pdf<br />

Zheldak, V. 2008. Development of Model <strong>Forest</strong> in Russia. Ustoychivoe lesopolzovanie, 2 (18),<br />

12-17 (In Russian).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P. Angelstam & M. Elbakidze. 2010. Sustainable forest management, multi-stakeholder governance <strong>and</strong> spatial planning<br />

544<br />

Table 1: Spatial scales for different types of spatial planning identified from interviews with 31 users of<br />

forest l<strong>and</strong>scapes’ goods, ecosystem services <strong>and</strong> values in the Kovdozersky Model <strong>Forest</strong> l<strong>and</strong>scape as<br />

determined by the stakeholders’ use profiles.<br />

Spatial scale Type of planning L<strong>and</strong>scape actor<br />

Trees in st<strong>and</strong>s<br />

Operational planning (e.g., <strong>Forest</strong> leasers that harvest wood<br />

(~1-100 ha)<br />

general considerations in<br />

forest management, stream<br />

<strong>and</strong> riparian management)<br />

St<strong>and</strong>s in management sub-unit<br />

(=l<strong>and</strong>scape, old lesnichestvo,<br />

leasing area)<br />

(~2,000 to 100,000 ha)<br />

<strong>L<strong>and</strong>scapes</strong> in a management<br />

unit (lesnichestvo under the<br />

2007 <strong>Forest</strong> Code)<br />

(~500,000 ha)<br />

River catchment in the boreal<br />

forest<br />

(~5,000,000 ha)<br />

Tactical planning (e.g., forest<br />

management, l<strong>and</strong>scape<br />

planning for game species)<br />

Strategic planning in the<br />

Kovdozersky MF (e.g., to<br />

assure co-existence of use of<br />

wood, non-wood goods,<br />

energy, tourism)<br />

Regional planning for<br />

sustainable development (e.g.,<br />

hydro power, tourism,<br />

conservation of ”green belt”<br />

forest)<br />

Kovdozersky forest management<br />

unit, nature protection units<br />

Kovdozersky forest management<br />

unit<br />

Hydroelectricity production,<br />

authorities working with<br />

nature conservation<br />

Hydropower station<br />

Kutsa zakaznik<br />

Research station<br />

of the MSU<br />

Gardening local societies<br />

Former sovhoz ”Khyazhegubsky”<br />

Mobile stations<br />

Zelenoborsky<br />

Lesozavodskoy<br />

Tourism enterprises<br />

Lesnichestvo<br />

Fish breeding factory<br />

Prison<br />

Zhyleks<br />

Murmanautodor<br />

Regiment OU<br />

<strong>Forest</strong> enterprises (7)<br />

Oktyabrskaya<br />

Railway road<br />

K<strong>and</strong>alaksha<br />

zapovednik<br />

Figure 1: Management interactions between l<strong>and</strong>scape stakeholders in the Kovdozersky Model <strong>Forest</strong> in<br />

the summer of 2008. Arrows show management interactions among different l<strong>and</strong>scape stakeholders.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Castro et al. 2010. Relationship between small ruminants behaviour <strong>and</strong> l<strong>and</strong>scape features in Northeast of Portugal<br />

545<br />

Relationship between small ruminants behaviour <strong>and</strong> l<strong>and</strong>scape<br />

features in Northeast of Portugal<br />

Marina Castro 1,2* , José Ferreira Castro 1 & António Gómez Sal 3<br />

1<br />

<strong>Escola</strong> <strong>Superior</strong> Agrária - Instituto Politécnico de Bragança, Campus de Santa<br />

Apolónia, Apartado 172, 5301-854 Bragança, Portugal<br />

2<br />

CIMO – Centro de Investigação de Montanha - Instituto Politécnico de Bragança,<br />

Portugal<br />

3<br />

Dpto. Interuniversitario de Ecología, Universidad de Alcalá. Edif. de Ciencias,<br />

Campus Universitario, 28871 Alcalá de Henares, Madrid, Spain<br />

Abstract<br />

The small ruminant production systems in Northeastern Portugal are mainly based on the<br />

extensive exploitation of the spontaneous plant production. The shepherds direct their flocks on<br />

daily grazing itineraries across different patches of l<strong>and</strong> use.<br />

Sheep <strong>and</strong> goats flocks were monitored monthly for a year. Data collected consists of<br />

geographical position <strong>and</strong> the type of l<strong>and</strong> use crossed. Also, essential livestock activities were<br />

monitored.<br />

The corrected frequencies (preference indexes) approach was used for the data analysis.<br />

The principal aims were to examine the relationships between livestock behaviour <strong>and</strong> l<strong>and</strong> use<br />

types, <strong>and</strong> to check how they change throughout the year <strong>and</strong> the time of day. Our results<br />

showed a strong dependence between l<strong>and</strong> use types <strong>and</strong> livestock activities <strong>and</strong> suggested a<br />

considerable coherence between human management, the spontaneous behaviour <strong>and</strong><br />

physiological needs of animals <strong>and</strong> the agroecosystems capacity to supply the livestock needs.<br />

Keywords: preference indexes, sheep <strong>and</strong> goat, livestock behaviour, l<strong>and</strong> use types<br />

1. Introduction<br />

The small ruminant production systems in Northeastern Portugal are mainly based on the<br />

extensive exploitation of the spontaneous plant production. The shepherds direct their flocks on<br />

daily grazing itineraries across different patches of l<strong>and</strong> use (Barbosa <strong>and</strong> Portela 2000; Castro<br />

et al. 2004). These circuits strongly differ throughout the year in duration <strong>and</strong> design. The<br />

places visited <strong>and</strong> the time spent in each one depends on the natural conditions <strong>and</strong> nutritional<br />

needs of the animals.<br />

The relationship between environmental factors <strong>and</strong> animal-behavior was studied by various<br />

authors (De Miguel et al. 1997; Oom et al. 2004; Horne et al. 2008). Most of them are focus on<br />

wild ungulate herbivorous or free-ranging domestic herbivorous. Livestock systems with human<br />

control such as itinerant shepherding haven’t been to a great extent studied from these point<br />

view. However, according (Baumont et al. 2000) shepherding consists in interacting with<br />

spontaneous animal’s decisions <strong>and</strong> the herder’s interventions could be considered simply as<br />

new constraints to the expression of the behavioral trends of the flock. As a result, in this feature,<br />

* Corresponding author. Tel.351 273 303345<br />

Email address: marina.castro@ipb.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Castro et al. 2010. Relationship between small ruminants behaviour <strong>and</strong> l<strong>and</strong>scape features in Northeast of Portugal<br />

546<br />

a study of livestock activities or movements across the l<strong>and</strong>scape permits underst<strong>and</strong>ing the<br />

animal’s perception of l<strong>and</strong>scape.<br />

Several animal attributes <strong>and</strong> environmental characteristics affect livestock movements, for<br />

instance, species or breed (Bailey et al. 2001), prior experience with a l<strong>and</strong>scape(Bailey et al.<br />

1996), degree of slope (Ganskopp <strong>and</strong> Vavra 1987), diurnal temperature dynamics of the<br />

l<strong>and</strong>scape(George et al. 2007), presence of trails (Ganskopp et al. 2000), resources availability<br />

<strong>and</strong> quality.<br />

Animal activities (grazing, resting <strong>and</strong> walking) were also affected by l<strong>and</strong>scape attributes<br />

(Ganskopp <strong>and</strong> Bohnert 2009). The research provided an underst<strong>and</strong>ing of how to animals use<br />

the l<strong>and</strong>scape, what kind of requirements they search when they visit a particular l<strong>and</strong> use type.<br />

Having an underst<strong>and</strong>ing of animal l<strong>and</strong>scape use could help to develop strategies to better<br />

management the l<strong>and</strong>scapes <strong>and</strong> its temporal changes.<br />

In this paper we show that relationships between animal behavior <strong>and</strong> environment can be easily<br />

highlighted by preference indexes values (Godron 1965). This method produces a general<br />

descriptive overview of the environmental factors associated with each type of behavior, <strong>and</strong><br />

provides information about the importance of each factor in conditioning animal activities.<br />

The principal aims were to examine the relationships between livestock behavior <strong>and</strong> l<strong>and</strong> use<br />

types, <strong>and</strong> to check how they change throughout the year <strong>and</strong> the time of day (temperature <strong>and</strong><br />

vegetation moisture effect). Our results showed a strong dependence between l<strong>and</strong> use types <strong>and</strong><br />

livestock activities <strong>and</strong> suggested a considerable coherence between human management, the<br />

spontaneous behavior <strong>and</strong> physiological needs of animals <strong>and</strong> the agroecosystems capacity to<br />

supply the livestock needs.<br />

2. Methodology<br />

The experiment was carried out in Trás-os-Montes region; from May 1999 to May 2000.<br />

Fieldwork was conducted over the territory of four villages located near Bragança, northeast<br />

Portugal (41°46’N latitude <strong>and</strong> 6°45’W longitude) at 700 to 1000 meters above sea level. The<br />

climate is humid Mediterranean with yearly mean temperature of 11.6°C <strong>and</strong> precipitation of<br />

972.1 mm, which occurs mainly from October until May (INMG 1991). The dominant soils are<br />

umbric leptosols <strong>and</strong> dystric leptosols, depending on the l<strong>and</strong> use.<br />

Four flocks (two of goat <strong>and</strong> two of sheep) were monitored every month for a year. Each flock<br />

was observed for a complete day by an operator using a GPS. Data collected consists of<br />

geographical position <strong>and</strong> type of l<strong>and</strong> use crossed (annual <strong>and</strong> perennial crops; meadows,<br />

forestl<strong>and</strong>s <strong>and</strong> scrubl<strong>and</strong>s). Also, animal behaviour was monitored.<br />

Behavioural activities (grazing, browsing, resting <strong>and</strong> walking) <strong>and</strong> the grazed species were<br />

noted every 15 minutes by direct observation (instantly recorded). Within each day, the<br />

frequency of animals involved in each activity was calculated for each individual observation of<br />

sheep or goat behaviour. The calculation of frequencies involved all the activities sampled,<br />

namely grazing, browsing, resting <strong>and</strong> walking.<br />

The corrected frequencies (preference indexes) approach was used for the data analysis. The<br />

frequency of sheep <strong>and</strong> goats in each activity was computed, in addition to each l<strong>and</strong> use type<br />

<strong>and</strong> part of the day. Also, the seasonal variations were computed of animal frequencies in each<br />

l<strong>and</strong> use in each time of the day (table 1).<br />

Table 1: Details of different frequencies considered<br />

Animals observed in each activity<br />

Animals observed in each l<strong>and</strong> use type<br />

* L<strong>and</strong> use type<br />

*Period of the year (cool days, warm days, summer,<br />

winter)<br />

*Part of day (morning, middle-day, afternoon) in different<br />

periods of year (cool days, warm days, winter, summer)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Castro et al. 2010. Relationship between small ruminants behaviour <strong>and</strong> l<strong>and</strong>scape features in Northeast of Portugal<br />

547<br />

The day was classified into morning, middle-day <strong>and</strong> afternoon, by dividing the daylight period.<br />

The year was classified into four periods; the summer corresponds to the months of June, July,<br />

August <strong>and</strong> September; cool days to the November, January, February, March; warm days to the<br />

April. May, October <strong>and</strong> winter December <strong>and</strong> January.<br />

Animal-environmental interactions were analysed by comparing the expected <strong>and</strong> observed<br />

frequencies (observed / expected). This approach shows the main patterns of association<br />

between animal activities <strong>and</strong> environmental variables. In addition, the relationship between<br />

physical variables <strong>and</strong> l<strong>and</strong> use types permits the analysis of the l<strong>and</strong>scape organisation <strong>and</strong> the<br />

animals ‘perception of the environment’.<br />

Numerical output allows us to identify the level of the relationship between variables: the<br />

association is positive or preferred when the quotient is higher than to 1.24; indifferent, when<br />

there is a value between 0.75 <strong>and</strong> 1.24, <strong>and</strong> negative or not preferred, when the corrected<br />

frequency value is lower than 0.75).<br />

3. Result<br />

The animals’ perception of the environment are focused on habitat types or l<strong>and</strong> uses types.<br />

Table 2 shows the preference index values for the four principal activities <strong>and</strong> five principal<br />

l<strong>and</strong> uses types (annual crops, perennial crops, meadows, scrubl<strong>and</strong>s <strong>and</strong> forestl<strong>and</strong>s). Stables<br />

<strong>and</strong> paths were also considered.<br />

Table 2: Preference index values of activities in different l<strong>and</strong> uses <strong>and</strong> stable <strong>and</strong> path<br />

Activities Annual Perennial Meadows Shrubs <strong>Forest</strong> Stable Path<br />

crops<br />

Sheep<br />

grazing 1.63 1.03 1.94 0.12 0.10 0.07 0.00<br />

resting 0.22 0.45 0.20 1.97 2.54 2.82 0.08<br />

walking 0.63 1.46 0.25 0.14 0.07 0.05 7.24<br />

browsing 1.35 2.61 0.13 3.66 0.82 0.00 0.37<br />

Goats<br />

grazing 3.21 0.62 2.36 0.73 0.13 0.00 0.16<br />

resting 0.37 1.23 0.62 0.26 1.69 4.11 0.21<br />

walking 0.48 0.49 0.40 0.88 0.34 0.22 4.05<br />

browsing 0.78 1.25 0.99 1.57 1.24 0.00 0.34<br />

The relationship between animal activities <strong>and</strong> l<strong>and</strong> use types permits the investigation of the<br />

animal habitat preferences <strong>and</strong> activity patterns.<br />

Grazing activities are positively related to annual crop l<strong>and</strong> use (1.63; 3.21) <strong>and</strong> meadows (1.94;<br />

2.36) in both kinds of flocks.<br />

Browsing activities are positively related to annual <strong>and</strong> perennial crop areas <strong>and</strong> scrubl<strong>and</strong>s in<br />

the case of sheep, whereas goat flocks concentrate browsing activities on scrubl<strong>and</strong>s (3.66).<br />

Resting activities are positively related to patches of forestl<strong>and</strong>s (2.54), scrubl<strong>and</strong>s (1.97) <strong>and</strong><br />

stables (2.82) in sheep flocks. Goat flocks rest in stables or forestl<strong>and</strong>s (1.69).<br />

Sheep <strong>and</strong> goat flocks use mainly the path to walk; nevertheless perennial crop l<strong>and</strong>s are used<br />

by sheep to move into other l<strong>and</strong> use types.<br />

Annual crop l<strong>and</strong> uses are used by sheep for grazing <strong>and</strong> browsing; activities of resting <strong>and</strong><br />

walking are not frequent in this l<strong>and</strong> use type. Goats use preferentially annual crops for grazing.<br />

These results suggest that some ancient areas of annual crops were converted into long-term<br />

fallow or have been ab<strong>and</strong>oned. For that reason, sheep flocks use this l<strong>and</strong> use type for grazing<br />

<strong>and</strong> browsing.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Castro et al. 2010. Relationship between small ruminants behaviour <strong>and</strong> l<strong>and</strong>scape features in Northeast of Portugal<br />

548<br />

Perennial crop l<strong>and</strong>s are used by sheep, mainly for moving up from some l<strong>and</strong> use types to<br />

others (1.46) <strong>and</strong> browsing (2.61). Goats use them to browse (1.25). There isn’t an association<br />

between perennial crop areas <strong>and</strong> resting activity (0.45 <strong>and</strong> 1.23 for sheep <strong>and</strong> goats,<br />

respectively). It relates to the organisation <strong>and</strong> specific function of each l<strong>and</strong> use type, for<br />

example, the flocks rest in forest l<strong>and</strong>s <strong>and</strong> not in chestnut orchards, despite their availability<br />

<strong>and</strong> good shade.<br />

<strong>Forest</strong>l<strong>and</strong>s are used by goats (1.24) for browsing <strong>and</strong> resting (1.69), whereas sheep use this<br />

l<strong>and</strong> use for resting (2.54). Scrubl<strong>and</strong>s are used to feed sheep (3.66) <strong>and</strong> goats (1.57); they are<br />

also places for resting by sheep (1.97). Meadows are used only to feed in both species.<br />

Table 3 shows the preference index values from the five principal l<strong>and</strong> use types (also stable <strong>and</strong><br />

paths) in different periods of the year (cool days, warm days, winter, summer) of sheep <strong>and</strong><br />

goats flocks.<br />

Table 3: Preference index values of l<strong>and</strong> uses (also stable <strong>and</strong> path) in different periods of the year.<br />

Sheep<br />

Goats<br />

L<strong>and</strong> uses cool warm winter summer cool warm winter summer<br />

annual 1.56 1.46 0.62 0.63 1.63 0.88 0.95 0.77<br />

perennial 1.50 1.44 2.04 0.40 0.23 1.48 3.39 0.36<br />

meadows 1.24 0.83 2.03 0.84 1.08 0.47 1.54 1.20<br />

shrubs 1.00 0.26 0.00 1.63 1.50 1.35 1.11 0.45<br />

forest 0.07 0.20 0.11 1.92 0.13 1.00 0.47 1.60<br />

path 0.97 1.04 1.87 0.83 0.98 0.64 1.49 1.14<br />

stable 0.04 0.96 0.00 1.50 0.04 0.91 0.00 1.85<br />

In the case of sheep flocks, cool days are positively related to annual (1.56) <strong>and</strong> perennial crops<br />

(1.50) <strong>and</strong> negatively related to forestl<strong>and</strong>s. Goat flocks on cool days prefer to use annual crops<br />

(1.63) <strong>and</strong> scrubl<strong>and</strong>s (1.50). On warm days, sheep show the same pattern (annual <strong>and</strong> perennial<br />

crops) <strong>and</strong> goats replace annual with perennial crops (1.48). In the winter period goats <strong>and</strong><br />

sheep show the same pattern, using for the most part perennial crop l<strong>and</strong>s <strong>and</strong> meadows. In<br />

summer, itinerant sheep are strongly connected with forest (1.92) <strong>and</strong> shrub (1.63) l<strong>and</strong> use<br />

types. Itinerant goats are positively connected with forest l<strong>and</strong>s (1.60) <strong>and</strong> negatively related to<br />

shrub (0.45) <strong>and</strong> perennial (0.36) l<strong>and</strong> use types.<br />

The relationship between animal activities <strong>and</strong> time of day, as well as period of year, permits the<br />

investigation of how the animals underst<strong>and</strong> their environmental constraints <strong>and</strong> opportunities.<br />

There is a general widespread pattern for both species (table 4). In the morning, animals<br />

preferentially walk <strong>and</strong> browse. The mid-day period is used for resting. The afternoon period is<br />

used for grazing.<br />

Resting in both species is connected with the higher temperatures of mid day, in all periods of<br />

the year. The resting activity in goats in winter is unreliable. Sheep flocks show a stronger<br />

reluctance for morning grazing on cool days <strong>and</strong> in winter. In the case of goats, only in summer<br />

periods they are reluctant to graze in the mornings. In summer, all activities excluding repose<br />

are negatively connected with mid-day time.<br />

4. Discussion<br />

The relationship between activity preference <strong>and</strong> time of the day, l<strong>and</strong> use type, <strong>and</strong> their<br />

seasonal variation suggest a complex pattern of use of l<strong>and</strong>scape by the flocks.<br />

During their daily itineraries, the flocks use different l<strong>and</strong> use types for different purposes; they<br />

can be used for browsing, grazing, resting, or walking. Also, these uses can change throughout<br />

the year <strong>and</strong> day. The results confirm those found by (Castro 2004) with different<br />

methodologies; in particular, using the time spent in each l<strong>and</strong> use type.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Castro et al. 2010. Relationship between small ruminants behaviour <strong>and</strong> l<strong>and</strong>scape features in Northeast of Portugal<br />

549<br />

The data indicate that flocks move over the l<strong>and</strong>scape with a special perception of benefits <strong>and</strong><br />

requirements. The displacement is guided by a complex interpretation of l<strong>and</strong> uses type profits<br />

(fodder, shade or transit between habitats), environmental constraints like temperature of midday,<br />

moisture of pasture in the morning, <strong>and</strong> l<strong>and</strong> use types occurring near villages.<br />

Table 4: Seasonal variation of preference index activity values in different periods of day<br />

Sheep<br />

Goats<br />

warm days<br />

Activities morning mid-day afternoon morning mid-day afternoon<br />

grazing 1.06 0.76 1.32 0.51 0.78 1.52<br />

resting 0.07 1.80 0.43 0.44 1.61 0.46<br />

walking 1.42 0.82 0.99 1.29 0.75 1.20<br />

browsing 3.20 0.57 0.21 1.23 0.92 1.00<br />

summer<br />

grazing 1.17 0.00 1.76 0.90 0.53 1.62<br />

resting 0.63 1.92 0.52 0.09 2.29 0.35<br />

walking 1.91 0.00 0.99 1.20 0.21 1.24<br />

browsing 1.99 0.00 0.91 1.35 0.12 1.02<br />

cool days<br />

grazing 0.59 0.73 1.25 0.34 0.91 1.55<br />

resting 0.64 2.09 0.33 0.00 1.90 0.60<br />

walking 1.72 0.86 0.97 1.35 0.60 1.24<br />

browsing 3.36 1.68 0.14 1.48 1.00 0.67<br />

winter<br />

grazing 0.33 1.02 1.17 0.37 1.03 1.33<br />

resting 0.00 1.87 0.66 0.81* 0.75* 1.44*<br />

walking 1.81 0.70 0.99 1.34 0.54 1.39<br />

browsing 2.13 0.99 0.70 1.41 1.36 0.30<br />

*not reliable, because only six animals were observed in this activity<br />

The seasonal variation of frequencies in different l<strong>and</strong> use types for sheep <strong>and</strong> goats suggest that<br />

the itineraries vary during the year, regarding different needs of animals <strong>and</strong> dissimilar<br />

opportunities for exploitation of resources. Shepherds recognize these different habits <strong>and</strong> use<br />

them accordingly to manage their production systems (Meuret 1996; Baumont et al. 2000).<br />

The occurrence of resting was markedly associated with the places in which shelter was<br />

provided <strong>and</strong> the period of the year where the temperature increased. Various authors (De<br />

Miguel et al. 1997) described the same pattern.<br />

The feed activities (grazing <strong>and</strong> browsing) are related in the case of sheep to annual <strong>and</strong><br />

perennial crops, meadows <strong>and</strong> shrubs. Frequently, the meadows are enclosed by hedgerows or<br />

splashed by scattered riparian trees. In the case of goats, they related to annual crops, meadows<br />

<strong>and</strong> scrubl<strong>and</strong>s. An interesting aspect is the different value of browsing activity in meadows for<br />

sheep <strong>and</strong> goats. In the first case, they are negatively related. In the case of goats, the value<br />

(0.99) shows that browsing activity can take place in meadows (presence of isolated trees or<br />

hedges); also interesting is the value of browsing activity on scrubl<strong>and</strong>s for sheep (3.66) <strong>and</strong><br />

goats (1.57), showing goats browsing a bit everywhere while sheep browse specially on<br />

scrubl<strong>and</strong>s. These results agree with those reported by authors that pointed out the different<br />

foraging styles between sheep <strong>and</strong> goats.<br />

The complex pattern of l<strong>and</strong>scape use by flocks described in this paper, show the importance of<br />

each l<strong>and</strong> use type for a particular animal activity. The natural complexity of Mediterranean<br />

l<strong>and</strong>scapes has been increased by traditional pastoral activity. Several structures of the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Castro et al. 2010. Relationship between small ruminants behaviour <strong>and</strong> l<strong>and</strong>scape features in Northeast of Portugal<br />

550<br />

l<strong>and</strong>scape, such us hedgerows, scattered trees on the arable fields, forage trees, have been<br />

preserved over time by their functional value. However, part of these structures is threatened by<br />

the ab<strong>and</strong>onment of agriculture <strong>and</strong> rural areas. Also, this specific pastoral system is strongly<br />

threatened. As a result, specific measures for its conservation should be taken as soon as<br />

possible, in accordance with European L<strong>and</strong>scape Convention assumptions.<br />

References<br />

Bailey D. W., Gross J. E., Laca A. E., Rittenhouse L. R., Coughernour M. B., Swift D. M. <strong>and</strong><br />

Sims P. L. 1996. Mechanisms that result in large herbivore grazing distribution patterns.<br />

J. Range Manage., 49 (5): 386-400.<br />

Bailey D. W., Kress D. D., Anderson D. C., Boss D. L. <strong>and</strong> Miller E. T. 2001. Relationship<br />

between terrain use <strong>and</strong> perfomance of beef cows grazing foothill rangel<strong>and</strong>. J. Anim.<br />

Sci., 79: 1883-1891.<br />

Barbosa J. C. <strong>and</strong> Portela J. 2000. O pastoreio de percurso no sistema de exploração de ovinos<br />

em Trás-os-Montes. Pages 95-116 in Montmuro - a última rota da transumância. A. D. P.<br />

A., Viseu.<br />

Baumont R., Prache S., Meuret M. <strong>and</strong> Mor<strong>and</strong>-Fehr P. 2000. How forage characteristics<br />

influence behaviour <strong>and</strong> intake in small ruminants: a review. Livestock Production<br />

Science, 64: 15-28.<br />

Castro M. 2004. Análisis de la Interacción vegetación-herbívoro en sistemas silvo-pastorales<br />

basados en Quercus pyrenaica. Tesis Doctoral. Universidad de Alcalá, Alcalá de Henares.<br />

Castro M., Castro J. F. <strong>and</strong> Gómez Sal A. 2004. L'utilisation du territoire par les petits<br />

ruminants dans la région de montagne de Trás-os-Montes, au Portugal. In: J. P. Dubeuf<br />

(Eds). Evolution des systemes de production ovine et caprine: Avenir des systems<br />

extensifs face aux changements de la societe. Zaragoza, CIHEAM / FAO / IZCS /<br />

CIRVAL: pp.249-254.<br />

De Miguel J. M., Rodríguez M. A. <strong>and</strong> Gómez Sal A. 1997. Determination of animal behaviorenvironment<br />

relation-ships by correspondence analysis. J. Range Manage., 50 (1): 85-93.<br />

Ganskopp D., Cruz R. <strong>and</strong> Johnson D. E. 2000. Least-effort pathways: a GIS analysis of<br />

livestock trails in rugged terrain. Applied Animal Behaviour Science, 68: 179-190.<br />

Ganskopp D. <strong>and</strong> Vavra M. 1987. Slope use by cattle, feral horses, deer, <strong>and</strong> bighorn sheep.<br />

Northwest Sci., 61: 74-81.<br />

Ganskopp D. C. <strong>and</strong> Bohnert D. W. 2009. L<strong>and</strong>scape nutritional patterns <strong>and</strong> cattle distribution<br />

in rangel<strong>and</strong> pastures. Applied Animal Behaviour Science, 116: 110-119.<br />

George M., Bailey D., Borman M., Ganskopp D., Surber G. <strong>and</strong> Harris N. 2007. Factors <strong>and</strong><br />

Practices that influence Livestock Distribution. University of California Agriculture <strong>and</strong><br />

Natural Resources Communication Services, Oakl<strong>and</strong>, CA.<br />

Godron M. 1965. Les principaux types de profils écologiques. CEPE/CNRS, Montpellier. 8 p.<br />

Horne J. S., Garton E. O. <strong>and</strong> Rachlow J. L. 2008. A synoptic model of animal space use:<br />

Simultaneous estimation of home range, habitat selection, <strong>and</strong> inter/intra-specific<br />

relationships. Ecological Modelling, 214: 338-348.<br />

INMG 1991. O clima de Portugal. Normas climatológicas da região de "Trás-os-Montes e Alto<br />

Douro" e "Beira Interior", correspondentes a 1951 - 1980. INMG, Lisboa.<br />

Meuret M. 1996. Organizing a grazing route to motivate intake on coarse resources. Ann.<br />

Zootech., 45 (Suppl.): 87-88.<br />

Oom S. P., Beecham J. A., Legg C. J. <strong>and</strong> Hester A. J. 2004. Foraging in a complex<br />

environment: from foraging strategies to emergent spatial properties. Ecological<br />

Complexity, 1: 299-327.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Elbakidze et al. 2010. Does forest certification contribute to boreal biodiversity conservation<br />

551<br />

Does forest certification contribute to boreal biodiversity<br />

conservation Swedish <strong>and</strong> Russian experiences<br />

Marine Elbakidze¹ * , Per Angelstam¹, Kjell Andersson¹, Mats Nordberg¹ & Yurij Pautov²<br />

¹ School for <strong>Forest</strong> Engineers, Faculty of <strong>Forest</strong> Sciences, Swedish University of<br />

Agricultural Sciences, SE-739 21 Skinnskatteberg, Sweden<br />

² Silver Taiga Foundation, P.O. Box 810 Syktyvkar, 167 000 Komi Republic, Russian<br />

Federation<br />

Abstract<br />

<strong>Forest</strong> Stewardship Council (FSC) is one of the leading forest certification schemes, which<br />

encourages sustainable forest management. While many studies have been done concerning the<br />

political <strong>and</strong> social outcomes of FSC, little is known about the contribution of certification for<br />

biodiversity conservation on the ground. We analysed the FSC st<strong>and</strong>ard content for biodiversity<br />

conservation at different spatial scales, <strong>and</strong> outcomes of FSC implementation for boreal<br />

biodiversity conservation on-the-ground using concrete forests management units in Russia <strong>and</strong><br />

Sweden. Focusing on state forest management units in both countries we evaluated the<br />

connectivity of set-aside forests by applying morphological spatial pattern analyses. The<br />

Russian st<strong>and</strong>ard included spatial scales from tree <strong>and</strong> st<strong>and</strong> to l<strong>and</strong>scape <strong>and</strong> ecoregion, while<br />

the Swedish st<strong>and</strong>ard focused on finer scales. Areas set-aside for FSC were similar in both<br />

countries, but formal protection in the Russian study area was three times higher than in Sweden.<br />

Swedish set-aside core areas were two orders of magnitude smaller, had much lower<br />

connectivity <strong>and</strong> were located in a fragmented forestl<strong>and</strong> holding. To conclude, the potential of<br />

FSC for biodiversity conservation depends on the amount of formal protection, the spatial<br />

configuration of forest management units, <strong>and</strong> the functionality of habitat networks. Thus the<br />

areas certified according has limited information contents.<br />

Keywords: structural <strong>and</strong> functional connectivity, Sveaskog Co, Komi Republic, formally <strong>and</strong><br />

informally protected areas<br />

1. Introduction<br />

At the beginning of the 1990s different forest certification schemes appeared in order to<br />

promote implementation of sustainable forest management (SFM) policies in both developed<br />

<strong>and</strong> developing countries (e.g., Gulbr<strong>and</strong>sen 2005). The <strong>Forest</strong> Stewardship Council (FSC) <strong>and</strong><br />

the Program for the Endorsement of <strong>Forest</strong> Certification schemes (PEFC) are the two leading<br />

forest certification schemes, <strong>and</strong> both aim to encourage SFM implementation at different levels<br />

from local to global (Cashore et al. 2003, 2005; Auld et al. 2008; Gullison 2003).<br />

In this paper we focus on the certified forests in the boreal biome on the European continent.<br />

The FSC scheme is the only international certification system with wide geographical coverage<br />

in boreal forests in Europe, <strong>and</strong> the vast majority of the FSC certified forests in Europe are<br />

located in the boreal ecoregion. Sweden <strong>and</strong> the Russian Federation have the largest areas of<br />

FSC-certified forest in Europe, <strong>and</strong> both use wood to create forest products sold on international<br />

markets where certification matters to customers.<br />

* Corresponding author. Tel.: +46(0)581-660433 E-mail address: marine.elbakidze@smsk.slu.se<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Elbakidze et al. 2010. Does forest certification contribute to boreal biodiversity conservation<br />

552<br />

Many studies have been made concerning the political <strong>and</strong> social outcomes of FSC (e.g.,<br />

Cashore et al. 2003, 2005; Gulbr<strong>and</strong>sen 2005). In spite of ecological issues being the main<br />

concern at the initiation of the FSC system, little is known about the contribution of certification<br />

for biodiversity conservation on-the-ground (Brawn et. al. 2001; Gulisson 2003; Gulbr<strong>and</strong>sen<br />

2005; Rametsteiner <strong>and</strong> Simula 2003). The aim of this paper is to analyse the potential of FSC<br />

certification in terms of st<strong>and</strong>ard content <strong>and</strong> outcomes for boreal biodiversity conservation onthe-ground<br />

using large concrete forest management units in Sweden <strong>and</strong> the Russian Federation<br />

as study areas. We first analyse <strong>and</strong> compare the biodiversity conservation requirements at<br />

different spatial scales in national FSC st<strong>and</strong>ards in Sweden <strong>and</strong> Russia. Secondly, focusing on<br />

two large state forest management units we evaluate the structural connectivity of forests set<br />

aside for biodiversity conservation by applying morphological spatial pattern analyses (Vogt et<br />

al. 2007 a,b). Finally, we discuss the potential of FSC certification for biodiversity conservation<br />

with different levels of ambition in managed boreal forests in Russia <strong>and</strong> Sweden.<br />

2. Methodological framework<br />

2.1. Study areas<br />

Our study areas were the Bergslagen management unit (hereafter Bergslagen) of Sveaskog Co<br />

in south-central Sweden <strong>and</strong> the Priluzje forest management unit (hereafter Priluzje) in the<br />

Komi Republic in the Russian Federation. Bergslagen (59º N, 16º E) encompasses a total area of<br />

563,629 ha of forest l<strong>and</strong> ownership, including water <strong>and</strong> mires. The forested area consists of<br />

many forest polygons distributed in an area exceeding 4,000,000 ha within 9 counties in southcentral<br />

Sweden. The main forest tree species are Norway spruce (Picea abies) <strong>and</strong> Scots pine<br />

(Pinus silvestris). <strong>Forest</strong>s with domination of birches <strong>and</strong> aspen in younger succession stages<br />

occupy less then 8% of the total forested l<strong>and</strong>. Priluzje (60ºN; 49º E) occupies 810,252 ha, <strong>and</strong><br />

it forms one contiguous block of forested l<strong>and</strong>. The main tree species are Norway spruce (Picea<br />

abies) <strong>and</strong> Scots pine (Pinus silvestris). <strong>Forest</strong>s with domination of birches (Betula spp.) <strong>and</strong><br />

aspen (Populus tremula) occupy almost 40% of the total forested l<strong>and</strong> as a consequence of<br />

previous large-scale disturbances by fire <strong>and</strong> logging. Priluzje still hosts pristine forests with<br />

natural dynamics, <strong>and</strong> consequently near-natural composition, structure <strong>and</strong> functions.<br />

2.2. Analyses of the FSC st<strong>and</strong>ard content on biodiversity considerations<br />

There are, at least, four levels of ambitions for biodiversity conservation (Angelstam et al. 2004),<br />

viz.: (1) species may be present, but not in viable populations; (2) viable populations may be<br />

present, but only those that are not specialised on natural forest structures or having large area<br />

requirements; (3) communities of all naturally occurring species of the representative<br />

ecosystems of an eco-region are present, but large scale disturbances <strong>and</strong> global change can<br />

threaten their ecological integrity, <strong>and</strong> (4) ecosystems <strong>and</strong> governance systems have adaptive<br />

capacity <strong>and</strong> form resilient social-ecological systems (=l<strong>and</strong>scapes).<br />

The level of ambition for biodiversity conservation is correlated with the spatial scale of forest<br />

management. There are, as a minimum, four relevant spatial scales: (1) Trees in a st<strong>and</strong> cover<br />

individual trees <strong>and</strong> groups of trees within a forest st<strong>and</strong> (Eriksson <strong>and</strong> Hammer 2006). On this<br />

spatial scale species with very small habitat requirement could be maintained for a limited time,<br />

thus satisfying relatively low ambitions in biodiversity conservation (the first level of ambition<br />

above). (2) St<strong>and</strong>s in a l<strong>and</strong>scape correspond to the scale of a delimited area of similar tree<br />

composition, age structure, diameter <strong>and</strong> height (Eriksson <strong>and</strong> Hammer 2006). This spatial scale<br />

matches the maintenance of species with small area requirements such as vascular plants, but<br />

not viable populations of fungi <strong>and</strong> lichens in the long term (i.e., the first <strong>and</strong> second levels of<br />

ambitions). (3) <strong>L<strong>and</strong>scapes</strong> in an eco-region large enough to accommodate the needs of species<br />

with relatively large area requirements <strong>and</strong> habitat patch dynamics for species that track certain<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Elbakidze et al. 2010. Does forest certification contribute to boreal biodiversity conservation<br />

553<br />

successional stages (i.e., the third level of ambition). (4) Finally, eco-regions on the global level<br />

encompass the spatial scale which allows maintenance of ecosystem composition, structures,<br />

<strong>and</strong> functions linked to natural processes (Angelstam et al., 2004; Cabarle et al. 2005) (the<br />

highest level of ambition).<br />

The Russian <strong>and</strong> Swedish FSC st<strong>and</strong>ards (Russian National FSC 2008, Swedish National FSC<br />

2010) contain criteria <strong>and</strong> indicators (C&I), concerning ecological, economic <strong>and</strong> social-cultural<br />

dimensions of SFM. We assessed the extent to which current criteria <strong>and</strong> indicators capture the<br />

four spatial scales of biodiversity conservation in both st<strong>and</strong>ards.<br />

2.3. Area proportion of set-aside forests<br />

The study areas contained formally (according to the national legislation) <strong>and</strong> informally<br />

(voluntary within the FSC <strong>and</strong> company policy frameworks) protected forests, which were set<br />

aside for biodiversity conservation. To estimate the area <strong>and</strong> area proportion of formally <strong>and</strong><br />

informally protected areas for biodiversity conservation we selected all forests which are<br />

protected from relevant digital spatially explicit data bases used by the forest companies in<br />

Priluzje <strong>and</strong> Bergslagen. Spatial analyses were made for st<strong>and</strong> types with different age <strong>and</strong> tree<br />

species. In Priluzje the latest forest inventory was done in 1992 <strong>and</strong> was updated in 2002. In<br />

Bergslagen the last forest inventory was done in 1990, <strong>and</strong> it was then updated yearly taking<br />

into account the forest management activities undertaken.<br />

2.4. Assessment of structural connectivity of forest habitats<br />

The spatial configuration of forests set aside for biodiversity conservation is important to satisfy<br />

the requirements of species with different levels of ambitions for biodiversity conservation.<br />

According to Taylor et al. (1993, 2006), there are two kinds of l<strong>and</strong>scape connectivity:<br />

structural <strong>and</strong> functional. Structural connectivity describes only physical relations among habitat<br />

patches <strong>and</strong> does not provide functional connectivity if corridors are not used by target species.<br />

Functional connectivity is species-oriented <strong>and</strong> increases when some change in the l<strong>and</strong>scape<br />

structure increases the degree of movement or flows of organisms through the l<strong>and</strong>scape.<br />

To assess structural connectivity created by the forests set aside for biodiversity conservation in<br />

our two case studies we used Morphological Spatial Pattern Analyses (MSPA) (Vogt et al.<br />

2007a, 2007b; Ostapowicz et al. 2008). Following (Vogt et al. 2007a, 2007b), we considered<br />

seven classes of forest pattern: core, islet, edge, perforation, loop, bridge <strong>and</strong> branch. The<br />

informally <strong>and</strong> formally protected forests were merged to create the focal forest polygons for<br />

assessment of structural connectivity of forest habitats. All focal forest polygons were classified<br />

according to their age class <strong>and</strong> tree species composition. The vector forests maps derived from<br />

GIS were converted into raster maps with 25-m pixels. According to Ostapowicz et al. (2008),<br />

the MSPA classes on output maps depend on the size parameter used in the morphological<br />

model. From a biological perspective several studies show that the edges affect the environment<br />

in forest patches. We used the results of studies presented in (Aune et. al. 2005) that indicates<br />

that in the boreal forests edge effects vary among species groups, but that they generally extend<br />

at least 25 m into the forest <strong>and</strong> greater than 50 m for some groups. Therefore, in MSPA<br />

processing we quantify connectivity in our study areas twice with edge width of 25 <strong>and</strong> 50 m.<br />

3. Results<br />

3.1. FSC requirements for biodiversity conservation in Russian <strong>and</strong> Swedish st<strong>and</strong>ards<br />

In the Russian FSC st<strong>and</strong>ard biodiversity considerations were included into 14 criteria with 48<br />

indicators, <strong>and</strong> in the Swedish FSC st<strong>and</strong>ard to 12 criteria with 55 indicators. Multiple<br />

indicators considered different aspects of biodiversity conservation. The C&I in both st<strong>and</strong>ards<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Elbakidze et al. 2010. Does forest certification contribute to boreal biodiversity conservation<br />

554<br />

were thus relevant for biodiversity conservation at different spatial scales. Some indicators<br />

referred to only one spatial scale of biodiversity considerations, while other corresponded to<br />

several spatial scales. Our analyses showed that in the Russian st<strong>and</strong>ard four spatial scales were<br />

reflected, <strong>and</strong> the majority of C&I corresponded to the scales of st<strong>and</strong>s in a l<strong>and</strong>scape <strong>and</strong><br />

l<strong>and</strong>scapes in an ecoregion. In the Swedish st<strong>and</strong>ard three spatial scales were considered, <strong>and</strong><br />

the majority of C&I were relevant to the scales of trees in a st<strong>and</strong> <strong>and</strong> st<strong>and</strong>s in a l<strong>and</strong>scape.<br />

3.2. Area proportion of set-aside forests<br />

In Bergslagen the formally protected forests included Natura 2000 sites with high conservation<br />

interest at the EU level, nature reserves created for the purpose of conserving biological<br />

diversity, protecting <strong>and</strong> preserving valuable natural environments or satisfying the need of<br />

areas for outdoor recreation; <strong>and</strong> nature of national interest with high natural or cultural value<br />

which must be protected from actions that may significantly damage the natural environment.<br />

The formally protected forests occupied around 4% of total forested area in Bergslagen. There<br />

were also several types of informally protected forests such as: (a) forests belonging to the<br />

management objective class NO which are left untouched in order to protect its high natural<br />

values; (b) forests of the NS class to be managed in a way to maintain a high nature values, <strong>and</strong><br />

(c) forests of the PF class where the aim is to combine wood production with set-aside of trees<br />

<strong>and</strong> small patches of some nature/culture value. The informally protected forests in Bergslagen<br />

covered almost 9% of the total forested area.<br />

In Priluzje formally protected forests fell into the following categories: (a) forests with water<br />

protective functions along rivers <strong>and</strong> streams; (b) forests along roads; (c) forests along fish<br />

spawning places; <strong>and</strong> (d) special protected areas. The formally protected forests amounted to<br />

almost 12% of the total forest area in Priluzje. The informally protected forests included pristine<br />

forests, <strong>and</strong> forests with high social <strong>and</strong> cultural values for local people, which are excluded<br />

from commercial use. Together these forests occupied approximately 10% of total forested area.<br />

3.3. Structural connectivity<br />

In Bergslagen pine, coniferous <strong>and</strong> spruce forests together occupied 84% of the total forested<br />

area set aside for biodiversity conservation. The majority of the forest pattern classes (core, edge,<br />

bridge <strong>and</strong> branch) were associated with these forest types, <strong>and</strong> the mostly with pine forests.<br />

Mixed <strong>and</strong> deciduous forests were underrepresented, <strong>and</strong> the main pattern classes created by<br />

these forests were edges <strong>and</strong> branches. By contrast, in Priluzje deciduous <strong>and</strong> mixed forests<br />

were the dominant forest types set aside for biodiversity conservation, <strong>and</strong> occupied 61% of the<br />

total area of formally <strong>and</strong> informally protected forests. More than half of the core areas were<br />

represented by deciduous <strong>and</strong> mixed forests. Edges were the second pattern class by the size of<br />

occupied area, <strong>and</strong> were represented mainly by deciduous <strong>and</strong> mixed forests.<br />

In Bergslagen core <strong>and</strong> edge were the two dominant classes in the forest pattern when the edge<br />

width was defined as 25 m. Taken together these two classes occupied 69% of forested area set<br />

aside for biodiversity conservation. When the edge width was increased to 50 m the forest<br />

pattern became very different. The area of core decreases almost three times (from 35 to 12%),<br />

<strong>and</strong> islet increased almost four times <strong>and</strong> became the single dominant pattern class. In Priluzje<br />

the forest pattern was very different. With an edge width of 25 m, core was the only dominant<br />

class, occupying 70% of the total area of forest set aside for biodiversity. The total area of two<br />

classes in Priluzje (core <strong>and</strong> edge) was equal to the sum of areas of four classes (core, edge,<br />

branch <strong>and</strong> islet) in Bergslagen. This could indicate that in Bergslagen the forests set aside for<br />

biodiversity were more fragmented than in Priluzje. With an edge width of 50 m the forest<br />

pattern changed. The core areas decreases to 47%, <strong>and</strong> the area of branch <strong>and</strong> edge increased.<br />

However, these changes in proportion of forest classes were much less pronounced than in<br />

Bergslagen.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Elbakidze et al. 2010. Does forest certification contribute to boreal biodiversity conservation<br />

555<br />

The total number of cores in the forest pattern of Bergslagen was 11,172 (with edge of 25 m)<br />

<strong>and</strong> 3,662 (with edge of 50 m). The size of cores ranged from 0.06 to 941 ha. The majority of<br />

cores (almost 70% of the total number) were less than 1 ha large. Core areas from 10 to 100 ha<br />

constituted only 6% of the total number core areas, but included more than 40% of the total core<br />

area. The total number of core areas in Priluzje was much smaller than in Bergslagen, <strong>and</strong><br />

amounted to 227 with 25-m edge <strong>and</strong> 207 with 50-m edge. The minimum size of a core was<br />

0.06 ha <strong>and</strong> the maximum was 37,397 ha. The majority of cores ranged from 0.1 to 10 ha.<br />

However, more than 90% of the total core area were more than 1,000 ha large.<br />

Based on analyses of the habitat selection of red-listed species (e.g., Berg et al. 1994) <strong>and</strong> the<br />

focal species approach (e.g., Lambeck 1997) applied to boreal focal species, we selected the<br />

most valuable cores of different forest types for focal species depending on old <strong>and</strong> old-growth<br />

forests. We selected cores of spruce, pine, coniferous, mixed <strong>and</strong> deciduous forests in the age of<br />

more than 110 years, which belong to the groups of old <strong>and</strong> old-growth forests in Bergslagen<br />

<strong>and</strong> Priluzje.<br />

There were 3,668 valuable cores in Priluzje <strong>and</strong> 4,940 cores in Bergslagen with an edge width<br />

of 25 m. Almost 80% of old <strong>and</strong> old-growth forests’ cores in Priluzje were a part of larger cores<br />

of forest area which were set aside for biodiversity conservation. The total number of valuable<br />

cores in Bergslagen decreases to 1,207 if the edge width equals to 50 m, <strong>and</strong> in Priluzje the<br />

same changes occur but not so extreme – down to 3,233 valuable cores. The total area of<br />

valuable cores in Priluzje was almost 6 times larger then in Bergslagen (39,413 ha in Priluzje<br />

<strong>and</strong> 6,299 ha in Sweden with the edge of 25 m). These differences increase with edge width of<br />

50 m. The area of valuable cores decreases in both study areas but more in Bergslagen (down to<br />

24,617 ha in Priluzje <strong>and</strong> to 2,289 ha in Bergslagen). These changes indicate that the edge size<br />

affects the number <strong>and</strong> area of valuable habitats in Bergslagen much stronger than in Priluzje.<br />

The sizes of valuable cores for biodiversity varied considerable in our study areas. For example,<br />

in Priluzje total number of cores was bigger in the size interval from 1 to 10 ha.However, the<br />

largest area of old <strong>and</strong> old-growth forests’ cores lay in the size interval from 10 to 100 ha. In<br />

Bergslagen the core distribution was different. The majority of cores were between 0.1 to 1 ha,<br />

<strong>and</strong> the most of the old <strong>and</strong> old-growth forest core areas were from 1 to 10 ha.<br />

4. Discussion<br />

Analysis of the content of the Russian <strong>and</strong> Swedish FSC st<strong>and</strong>ards showed that the Russian<br />

st<strong>and</strong>ard contained higher biodiversity conservation ambitions, thus including maintenance of<br />

communities of all naturally occurring species of the representative ecosystems of an eco-region.<br />

By contrast the C&I of the Swedish FSC st<strong>and</strong>ard were more focused on the maintenance of<br />

species, which are not specialised on natural forest structures or have large area requirements.<br />

The total area of formally <strong>and</strong> informally protected areas as a proportion of the total forested<br />

area of the forest management units in Priluzje was almost 70% larger than in Bergslagen, <strong>and</strong><br />

approximately half of these forests in Priluzje were set aside according to the national<br />

legislation. In Bergslagen the area of voluntary protected forests was almost twice as large as<br />

the area of formally protected forests. The large difference in the spatial configuration of the<br />

forest holdings in the Bergslagen case study (a dispersed archipelago of forest holdings) <strong>and</strong> the<br />

Priluzje case study (one contiguous patch) should be noted.<br />

A review of the patch size requirements of individuals of different groups of species indicate<br />

that the core patch size distribution was satisfactory in the Swedish case study mainly for plants,<br />

fungi <strong>and</strong> lichens, but not for birds <strong>and</strong> mammals. By contrast, the majority of core areas in the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M. Elbakidze et al. 2010. Does forest certification contribute to boreal biodiversity conservation<br />

556<br />

Russian case study had the potential to host viable populations also of the most area-dem<strong>and</strong>ing<br />

focal species.<br />

To conclude, biologically relevant analyses are needed to assess functionality of set-asides for<br />

biodiversity conservation. We conclude that the potential of FSC for biodiversity conservation<br />

is dependent both on the amount of formal protection, the spatial configuration of forest<br />

management units, <strong>and</strong> the functionality <strong>and</strong> renewal of habitat networks.<br />

References<br />

Auld, G., Gulbr<strong>and</strong>sen, L. <strong>and</strong> McDermott, C. 2008. Certification schemes <strong>and</strong> the impacts of<br />

forests <strong>and</strong> forestry. Annual Review of Environment <strong>and</strong> Resources, 33: 9.1 – 9.25.<br />

Aune, K., Jonsson, B.-J. <strong>and</strong> Moen, J. 2005. Isolation <strong>and</strong> edge effects among woodl<strong>and</strong> key<br />

habitats in Sweden: Is forest policy promoting fragmentation Biological conservation, 124:<br />

89-95.<br />

Brawn, N., Noss, R., Diamond, D. <strong>and</strong> M., Myers. 2001. Conservation biology <strong>and</strong> forest<br />

certification: working together toward ecological sustainability. Journal of <strong>Forest</strong>ry, 18-25.<br />

Cashore, B., Auld, G. <strong>and</strong> Newsom, D. 2003. <strong>Forest</strong> certification (eco-labeling) programs <strong>and</strong><br />

their policy-making authority: explaining divergence among North America <strong>and</strong> European<br />

case studies. <strong>Forest</strong> policy <strong>and</strong> economics, 5: 225-247.<br />

Cashore, B., van Kooten, C., Vertinsky, I., Auld, G. <strong>and</strong> Affolderbach, J. 2005. Private or selfregulation<br />

A comparative study of forest certification choices in Canada, the United States<br />

<strong>and</strong> Germany. <strong>Forest</strong> policy <strong>and</strong> economics, 7: 53-69.<br />

Eriksson, S. <strong>and</strong> Hammer, M. 2006. The challenge of combining timber production <strong>and</strong><br />

biodiversity conservation for long-term ecosystem functioning – a case study of Swedish<br />

boreal forestry. <strong>Forest</strong> ecology <strong>and</strong> management, 237: 208 -217.<br />

Esseen, P.-A. <strong>and</strong> Renhorn, K.-E., 1998. Edge effects on an epiphytic lichen in fragmented<br />

forests. Conservational Biology, 12 6: 1307–1317<br />

Gulbr<strong>and</strong>sen, L. 2005. The effectiveness of non-state governance schemes: a comparative study<br />

of forest certification in Norway <strong>and</strong> Sweden. International Environmental Agreements, 5:<br />

125-149.<br />

Gulisson, R.F. 2003. Does forest certification conserve biodiversity Oryx, 37 (2): 154-165<br />

Lambeck, R.J. 1997. Focal species: a multi-species umbrella for nature conservation.<br />

Conservation Biology, 11 (4): 849-856.<br />

Ostapowicz, K., Vogt, P., Ritters, K., Kozak, J. <strong>and</strong> Estreguil, C. 2008. Impact of scale on<br />

morphological spatial pattern of forest. L<strong>and</strong>scape ecology, 23: 1107-1117.<br />

Rametsteiner, E. <strong>and</strong> Simula, M. 2003. <strong>Forest</strong> certification – an instrument to promote<br />

sustainable forest management Journal of environmental management, 67: 87-98.<br />

Russian National <strong>Forest</strong> Stewardship st<strong>and</strong>ard. 2008.<br />

Swedish FSC st<strong>and</strong>ard for forest certification. 2010.<br />

Taylor P., Fahrig, L. <strong>and</strong> With,K. 2006. L<strong>and</strong>scape connectivity: a return to the basics. In:<br />

Crooks, K. <strong>and</strong> Sanjayan, M. (Eds). Connectivity conservation. Cambridge University<br />

Press: 29 - 43.<br />

Taylor, P. <strong>and</strong> Fahrig, L., Henein, K., Merriam, G. 1993. Connectivity is a vital element of<br />

l<strong>and</strong>scape structure. Oikis, 68 (3): 571-572.<br />

Vogt, P., Riitters, K., Estreguil, C., Kozak, J., Wade, T. <strong>and</strong> Wickham, J. 2007. Mapping spatial<br />

patterns with morphological image processing. L<strong>and</strong>scape ecology, 22:171-177.<br />

Vogt, P., Riitters, K., Iwanowski, M., Estreguil, C., Kozak, J. <strong>and</strong> Soille, P. 2007. Mapping<br />

l<strong>and</strong>scape corridors. Ecological Indicators, 7 (2): 481-488.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Fonseca & T. de Figueiredo 2010. Impact of tree species replacement on carbon stocks in forest floor <strong>and</strong> mineral soil<br />

557<br />

Impact of tree species replacement on carbon stocks in forest floor <strong>and</strong><br />

mineral soil<br />

Felícia Fonseca * & Tomás de Figueiredo<br />

Centro de Investigação de Montanha (CIMO), Instituto Politécnico de Bragança<br />

(<strong>ESA</strong>B/IPB), Apartado 1172, 5301-855 Bragança, Portugal<br />

Abstract<br />

This study aims at evaluating the influence of replacing areas of Quercus pyrenaica, which<br />

represents native vegetation of Serra da Nogueira, NE of Portugal, by Pseudotsuga menziesii on<br />

carbon stocks in forest floor <strong>and</strong> mineral soil. Three sampling areas were selected in adjacent<br />

locations with similar soil <strong>and</strong> climate conditions. The first area, covered by Quercus pyrenaica<br />

(QP), represents the original soil. The second area is in a 40 years old st<strong>and</strong> of Pseudotsuga<br />

menziesii (PM40), <strong>and</strong> the third one, also under Pseudotsuga menziesii, is 15 years old (PM15).<br />

In each sampling area, at 10 r<strong>and</strong>omly selected points, samples were collected in the forest floor<br />

(0.49 m 2 quadrat) <strong>and</strong> in the soil (at 0-5, 5-10 <strong>and</strong> 10-20 cm depth). The forest floor stores 17,<br />

13 <strong>and</strong> 6% of total carbon for PM40, PM15 <strong>and</strong> QP, respectively. Four decades after species<br />

replacement, a soil organic carbon loss is observed, although no significant differences were<br />

found when comparing soil under introduced (PM) with original species (QP). A carbon loss of<br />

around 30%, in PM15, <strong>and</strong> gains of about 10%, in PM40, are computed when considering<br />

mineral soil <strong>and</strong> forest floor together.<br />

Keywords: Quercus pyrenaica; Pseudotsuga menziesii; forest floor; soil organic carbon.<br />

1. Introduction<br />

The increase in atmospheric carbon content, as expected considering actual trends, draws<br />

attention to the highly valuable role of forest ecosystems in the global carbon cycle (Eswaran et<br />

al., 1993). The majority of the native vegetation in Iberian peninsula (Portugal <strong>and</strong> Spain) have<br />

been replaced by other forest species, particularly fast-growing conifers plantations. Although<br />

this replacement may have beneficial economic consequences, it is essential to underst<strong>and</strong><br />

environmental effects such as in carbon sequestration for mitigation of greenhouse gases. The<br />

knowledge of the differences among species in what regards C sequestration should be a<br />

decision support tool when introducing new forest species <strong>and</strong> can be used strategically to reach<br />

environmental goals (Oostra et al., 2006; Schulp et al., 2008; Vallet et al., 2009).<br />

Species replacement implies changes in carbon stocks in forest floor <strong>and</strong> soil organic<br />

matter (Peltoniemi et al., 2004) because tree species litter quantity, quality <strong>and</strong> distribution in<br />

soil horizons have high influence in carbon storage (Oostra et al., 2006). Decomposition rate of<br />

plant residues can be slower or faster, depending on their nature. In general, it is accepted that<br />

organic residues from coniferous species decompose more slowly than broadleve species, for<br />

example, due to the presence of non-hydrolyzable polyphenolic compounds in litter (Faulds <strong>and</strong><br />

Williamson, 1999). On the other h<strong>and</strong>, fast-growing species would accumulate carbon more<br />

rapidly than slow-growing species, but several studies shown that substitution leads to a carbon<br />

* Corresponding author.<br />

Email address: ffonseca@ipb.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Fonseca & T. de Figueiredo 2010. Impact of tree species replacement on carbon stocks in forest floor <strong>and</strong> mineral soil<br />

558<br />

loss (Schroth et al., 2002; Wang <strong>and</strong> Wang, 2007; Vallet et al., 2009). Carbon stored in forest<br />

ecosystems depends fundamentally on forest age <strong>and</strong> management pratices (Post <strong>and</strong> Kwon,<br />

2000; Paul et al., 2002; Pregitzer <strong>and</strong> Euskirchen, 2004). Species choice is actually a<br />

management option to increase carbon storage (Vallet et al., 2009).<br />

The main objective of the present study was to quantify the impact of replacing a native<br />

hardwood species (Quercus pyrenaica) by a fast-growing conifer plantation (Pseudotsuga<br />

menziesii) on carbon stocks in forest floor <strong>and</strong> mineral soil, <strong>and</strong> its persistence through time.<br />

2. Methodology<br />

The study area was located in Serra da Nogueira, northeast of Portugal (41º 45'N <strong>and</strong> 6°<br />

52'W), in the range between 1000 <strong>and</strong> 1100 m altitude. The annual average temperature is 12º C<br />

<strong>and</strong> annual average precipitation is 1100 mm, concentrated from October to March (INMG,<br />

1991). The native vegetation is Quercus pyrenaica (QP), which occupies about 6,000 ha <strong>and</strong><br />

represents the area of QP most extensive in Portugal. Over the last decades, some of the QP<br />

area have been replaced by fast-growing species, mainly Pseudotsuga menziesii, process where<br />

the fires have had an important role. Soils are classified as Orthi-Dystric Leptosols derived of<br />

schists (Agroconsultores <strong>and</strong> Coba, 1991).<br />

To assess the impact of species replacement on carbon stocks in forest floor <strong>and</strong> mineral<br />

soil three sampling areas were selected in adjacent locations with similar soil <strong>and</strong> climate<br />

conditions. Selected study species were 15 year old (PM15) <strong>and</strong> 40 year old (PM40) st<strong>and</strong>s of<br />

Pseudotsuga menziesii <strong>and</strong> st<strong>and</strong>s of QP, the latter representing the original soil. The st<strong>and</strong>s<br />

characterization is presented in Table 1.<br />

St<strong>and</strong>s<br />

Table 1: Mean st<strong>and</strong> characteristics for Pseudotsuga menziesii <strong>and</strong> Quercus pyrenaica.<br />

Number of stems<br />

Age<br />

Dominant height<br />

Mean diameter<br />

Basal area<br />

(trees ha -1 ) (years) (m)<br />

(cm) (m 2 ha -1 )<br />

P. menziesii (PM40) 3800 40 20.1 27.7 229.3<br />

P. menziesii (PM15) 6700 15 13,9 16.6 144.3<br />

Q. pyrenaica (QP) 19300 - 8.5 7.3 81.6<br />

In each one of the three st<strong>and</strong>s (PM15, PM40, QP) 10 areas of 70 x 70 cm (0,49 m 2 ) were<br />

r<strong>and</strong>omly established. In each one of these, the forest floor, defined as organic material<br />

deposited over mineral soil, was collected. <strong>Forest</strong> floor samples were dried at 65 ºC for 72 h to<br />

determine dry mass.<br />

Total soil organic C down to 20 cm depth (sampled in the same points where forest floor<br />

had been removed) was calculated from C concentrations determined in samples collected in the<br />

0-5, 5-10 <strong>and</strong> 10-20 cm soil layers. Bulk density (BD) was estimated at the same depths using<br />

the equation:<br />

BD = 100 / (%OM / BD OM ) + ((100 – %OM) / BD min soil ) (1)<br />

where OM is soil organic matter, BD OM the bulk density of organic matter (assumed to be<br />

0.244), BD min soil the mineral soil bulk density (assumed to be 1.64) (Post <strong>and</strong> Kwon, 2000; Paul<br />

et al., 2002).<br />

Samples for soil C were air dried <strong>and</strong> sieved to determine the coarse fraction (> 2 mm).<br />

All samples of forest floor <strong>and</strong> mineral soil were analyzed for total C by dry combustion (ISO,<br />

1994). Soil samples were tested with an acid-drop but no carbonates were detected, thus the<br />

total soil C was assumed to be comparable to soil organic C. <strong>Forest</strong> floor mass values were<br />

converted to carbon (t C ha -1 ) multiplying these values by the C concentration in dry matter.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Fonseca & T. de Figueiredo 2010. Impact of tree species replacement on carbon stocks in forest floor <strong>and</strong> mineral soil<br />

559<br />

Soil organic carbon (SOC, t C ha -1 ) was calculated multiplying C concentration by bulk density<br />

<strong>and</strong> thickness of the mineral soil layer with a correction for coarse elements content.<br />

3. Results <strong>and</strong> discussion<br />

Carbon concentration is significantly higher in forest floor under native species (QP), but<br />

the amount of organic residues accumulated on the soil surface is higher under the introduced<br />

species (PM40 <strong>and</strong> PM15). The carbon stocks under PM40 is about three times higher than the<br />

original soil, following the pattern PM40 > PM15 > QP (Figure 1). These results seem to be<br />

related to the decomposition rate, since under coniferous it is evident, on the surface, the<br />

presence of a large amount of slightly decomposed organic debris, unlike under the deciduous<br />

there is less amount of organic material, which suggests a more rapid decomposition <strong>and</strong><br />

subsequent connection to the mineral fraction. Similar results were obtained by Rapp (1984)<br />

<strong>and</strong> Fonseca (1999). Among the conifers, differences also appear to be due to the age of the<br />

st<strong>and</strong>s <strong>and</strong> the increase in density of canopy cover observed from PM15 to PM40. In temperate<br />

forests, forest floor remain relatively constant or increase with age, reaching a peak after about<br />

70 years of st<strong>and</strong> development (Pregitzer <strong>and</strong> Euskirchen, 2004).<br />

Figure 1: <strong>Forest</strong> floor mass (t ha -1 ), C concentration (%) <strong>and</strong> C stocks (t ha -1 ) in forest floor. For each<br />

variable, averages with the same letter are not significantly different (P


F. Fonseca & T. de Figueiredo 2010. Impact of tree species replacement on carbon stocks in forest floor <strong>and</strong> mineral soil<br />

560<br />

The 0-20 cm soil profile contained an average of 55.9, 82.1 <strong>and</strong> 86.3 t C ha -1 in the PM15,<br />

PM40 <strong>and</strong> QP, respectively (Figure 2). The proportion of soil organic carbon storage in relation<br />

to total (forest floor plus mineral soil 0-20 cm) was 87.5% for PM15, 84.5% for PM40 <strong>and</strong><br />

94.6% for QP. All soil layers show significantly lower carbon storage values under PM15, than<br />

under PM40 or QP. Despite the statistically non-significant differences found between PM40<br />

<strong>and</strong> QP in soil layers, after 40 years of species replacement there are still carbon losses in<br />

PM40, when compared to native vegetation (Figure 3). The differences observed in soils under<br />

PM15 are mainly due to the shorter recovery time since disturbance caused during st<strong>and</strong><br />

installation (Dick et al., 1998).<br />

Figure 2: Total forest floor <strong>and</strong> soil organic carbon storage (numbers above bars in t C ha -1 ) according to<br />

species. Numbers below bars indicate total soil organic carbon storage in t C ha -1 . For each layer,<br />

averages with the same letter are not significantly different (P


F. Fonseca & T. de Figueiredo 2010. Impact of tree species replacement on carbon stocks in forest floor <strong>and</strong> mineral soil<br />

561<br />

The replacement of native vegetation (QP) by a fast-growing specie (PM) has effects on<br />

C stocks that depend on the time scale considered. The additional C storage of the PM40 st<strong>and</strong>s<br />

compared to QP is 5.3 t C ha -1 . After 15 years (PM15) the C loss reaches 27.8 t C ha -1 . There are<br />

many factors <strong>and</strong> processes that determine the distribution <strong>and</strong> rate of change in soil organic<br />

carbon storage when the type of vegetation <strong>and</strong> soil management practices are changed, among<br />

which st<strong>and</strong>s age may assume great importance, as shown in this study.<br />

References<br />

Agroconsultores <strong>and</strong> Coba, 1991. Carta dos Solos do Nordeste de Portugal. UTAD, Vila Real.<br />

Alcázar, J., Rothwell, R.L., Woodard, P.M., 2002. Soil disturbance <strong>and</strong> the potential for erosion<br />

after mechanical site preparation. North. J. Appl. For., 19: 5-13.<br />

Dick, W.A., Blevins, R.L., Frye, W.W.m Peters, S.E., Christenson, D.R., Pierce, F.J., Vitosh,<br />

M.L., 1998. Impacts of agricultural management practices on C sequestration in forestderived<br />

soils of the eastern Corn Belt. Soil & Till. Tes., 47: 235-244.<br />

Eswaran, H., Berg, E.V.D., Reich, P., 1993. Organic carbon in soils of the World. Soil Sci. Soc.<br />

Am. J., 57: 192–194.<br />

Faulds, C.B., Williamson, G., 1999. The role of hydroxycinnamates in the plant cell wall. J. Sci.<br />

Food Agri., 79: 393–395.<br />

Fonseca, F., Martins, A., 1999. Interacções vegetação-solo em ecossistemas florestais no N. de<br />

Portugal: Natureza dos horizontes orgânicos e implicações no solo. Actas das<br />

Comunicações do XIV Congreso Latinoamericano de la Ciencia del Suelo, Universidad<br />

de la Frontera, Pucón, Chile. 6 pp.<br />

INMG, 1991. Normais Climatológicas da Região de "Trás-os-Montes e Alto Douro" e "Beira<br />

Interior" Correspondentes a 1951-1980. Fascículo XLIX, Volume 3, 3ª Região, Lisboa.<br />

ISO, 1994. Organic <strong>and</strong> total carbon after dry combustion. In: Environment soil quality.<br />

ISO/DIS 10694.<br />

Oostra, S., Majdi, H., Olsson, M., 2006. Impact of tree species on soil carbon stocks <strong>and</strong> soil<br />

acidity in southern Sweden. Sc<strong>and</strong>inavian Journal of <strong>Forest</strong> Research, 21: 364–371.<br />

Paul, K.I., Polglase, P.J., Nyakuengama, J.G., Khanna, P.K., 2002. <strong>Change</strong> in soil carbon<br />

following afforestation. <strong>Forest</strong> Ecology <strong>and</strong> Management, 168: 241–257.<br />

Peltoniemi, M., Mäkipää, R., Liski, J., Tamminen, P., 2004. <strong>Change</strong>s in soil carbon st<strong>and</strong> age –<br />

an evaluation of a modeling method with empirical data. <strong>Global</strong> <strong>Change</strong> Biology, 10:<br />

2078-2091.<br />

Pregitzer, K., Euskirchen, E., 2004. Carbon cycling <strong>and</strong> storage in world forests: biome patterns<br />

related to forest age. <strong>Global</strong> <strong>Change</strong> Biology, 10: 2052–2077.<br />

Post, W.M., Kwon, K.C., 2000. Soil carbon sequestration <strong>and</strong> l<strong>and</strong>-use change: processes <strong>and</strong><br />

potential. <strong>Global</strong> <strong>Change</strong> Biology, 6: 317–327.<br />

Rapp, M., 1984. Répartition et flux de matière organique dans un écosystème à Pinus pinea L.<br />

Ann. Sci. For. 41: 253-272.<br />

Schroth, G., D’Angelo, S.A., Teixeira, W.G., Haag, D., Lieberei, R., 2002. Conversion of<br />

secondary forest into agroforestry <strong>and</strong> monoculture plantations in Amazonia:<br />

consequences for biomass, litter <strong>and</strong> soil carbon stocks after 7 years. Ecology <strong>and</strong><br />

Management, 163: 131–150.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


F. Fonseca & T. de Figueiredo 2010. Impact of tree species replacement on carbon stocks in forest floor <strong>and</strong> mineral soil<br />

562<br />

Schulp, C.J.E., Nabuurs, G-J., Verburg, P.H., de Waal, R.W., 2008. Effect of tree species on<br />

carbon stocks in forest floor <strong>and</strong> mineral soil <strong>and</strong> implications for soil carbon inventories.<br />

<strong>Forest</strong> Ecology <strong>and</strong> Management, 256: 482–490.<br />

Vallet, P., Meredieu, C., Seynave, I., Bélouard, T., Dhôte, J.F., 2009. Species substitution for<br />

carbon storage: Sessile oak versus Corsican pine in France as a case study. <strong>Forest</strong><br />

Ecology <strong>and</strong> Management, 257: 1314–1323.<br />

Wang, Q.K., Wang, S.L., 2007. Soil organic matter under different forest types in Southern<br />

China. Geoderma, 142: 349–356.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Gil-Tena et al. 2010. Disentangling recent changes in forest bird ranges in Mediterranean forests (NE Spain)<br />

563<br />

Disentangling recent changes in forest bird ranges in Mediterranean<br />

forests (NE Spain): Assessing global change impacts <strong>and</strong> guiding<br />

l<strong>and</strong>scape management<br />

Assu Gil-Tena 1, 2* , Lluís Brotons 2, 3 & Santiago Saura 4<br />

1 Departament d’Enginyeria Agroforestal, ETSEA, Universitat de Lleida, Spain<br />

2 Àrea de Biodiversitat, Centre Tecnològic <strong>Forest</strong>al de Catalunya, Spain<br />

3 Institut Català d’Ornitologia, Museu de Ciències Naturals, Zoologia, Spain<br />

4 Departamento de Economía y Gestión <strong>Forest</strong>al, E.T.S.I. Montes, Universidad<br />

Politécnica de Madrid, Spain<br />

Abstract<br />

In Mediterranean Europe, after the widespread afforestation <strong>and</strong> forest maturation following a<br />

large-scale decline in traditional uses, many forest bird species have exp<strong>and</strong>ed their ranges in<br />

the last decades of the 20th century. In order to disentangle the processes mediating recent range<br />

changes of forest birds in the Mediterranean region of Catalonia (NE Spain) <strong>and</strong> to provide<br />

forest management guidelines, we studied the relationships between variations in forest bird<br />

species richness at 10x10 km <strong>and</strong> forest l<strong>and</strong>scape dynamics associated with l<strong>and</strong> ab<strong>and</strong>onment<br />

(afforestation <strong>and</strong> maturation), fires <strong>and</strong> management. The widespread afforestation <strong>and</strong> forest<br />

maturation appeared to counteract the potentially negative effects of fires on species richness,<br />

being the impact of forest management on birds much smaller than the impact of the former.<br />

Nevertheless, forest practices of moderate intensity <strong>and</strong> an adequate management of l<strong>and</strong>scape<br />

connectivity pattern may be beneficial, allowing species to better face range changes associated<br />

with global change.<br />

Keywords: afforestation, fires, forest maturation, species richness variation, sylvicultural<br />

treatments<br />

1. Introduction<br />

Ecosystems of Mediterranean Europe appear to be especially susceptible to the impacts of<br />

global change (Metzger et al. 2008) since many large-scale factors such as climate <strong>and</strong> l<strong>and</strong>-use<br />

changes or modifications in the perturbation regime are expected to simultaneously impact these<br />

regions, with largely unknown overall effects on current biodiversity patterns (Sala et al. 2000).<br />

However, forest management could help to buffer the expected negative impacts of global<br />

change in the years to come. For instance, forest management could mitigate climate change by<br />

modifying wildfire behaviour (De Dios et al. 2007). Therefore, providing guidelines for a<br />

proactive management is of fundamental in ecological research.<br />

As opposed to the current biodiversity crisis due to global change, many forest birds in<br />

Catalonia (NE Spain) have exp<strong>and</strong>ed or maintained their ranges in the last years of the 20th<br />

century (Estrada et al. 2004). Particularly, in this Mediterranean European region, l<strong>and</strong><br />

ab<strong>and</strong>onment has boosted afforestation (Poyatos et al. 2003) <strong>and</strong> forests have also significantly<br />

* Corresponding author. Tel.: +34 973702876 - Fax: +34 973702673<br />

Email address: agil@eagrof.udl.cat<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Gil-Tena et al. 2010. Disentangling recent changes in forest bird ranges in Mediterranean forests (NE Spain)<br />

564<br />

aged (Spanish National <strong>Forest</strong> Inventory; Ministerio de Medio Ambiente 1997-2007), although<br />

fires burnt approximately 240,000 ha between 1975 <strong>and</strong> 1998 (Díaz-Delgado et al. 2004).<br />

In order to provide effective forest management guidelines to better face global change impacts<br />

in Mediterranean Europe, this study aimed at determining the influence of forest l<strong>and</strong>scape<br />

dynamics at 10x10 km associated with l<strong>and</strong> ab<strong>and</strong>onment (afforestation <strong>and</strong> maturation), fires<br />

<strong>and</strong> management on the variation of forest bird species richness in Catalonia in the two last<br />

decades of the 20th century. For this purpose, we considered data from bird atlases, forest<br />

inventories, fire perimeters’ <strong>and</strong> l<strong>and</strong>-use maps.<br />

2. Methodology<br />

2.1 <strong>Forest</strong> bird changes<br />

Data on changes in the ranges of forest birds were gathered from the Catalan Breeding Bird<br />

Atlases (CBBA; Estrada et al. 2004). The CBBA data are derived from a series of large-scale<br />

surveys covering the whole extent of Catalonia in two different periods: 1975–1983 (Atlas1)<br />

<strong>and</strong> 1999–2002 (Atlas2). A total of 385 10x10 km UTM squares were surveyed in each of the<br />

different time periods. The field work was conducted by volunteers from March to July,<br />

recording evidence of breeding either by sound or sight.<br />

We considered the variation of species richness between atlases from the species occurrence<br />

data for each Atlas period. The analyzed species did not include: (1) the most common species<br />

(>90% of total squares in Atlas 2) or very scarce (7.5 cm) that had been inventoried in the NFI2 but that were not present in the<br />

same permanent plots in the NFI3. We considered only the silvicultural treatments that can<br />

affect st<strong>and</strong> basal area, differentiating between regeneration (clearcutting, shelterwood <strong>and</strong><br />

selective cutting) <strong>and</strong> st<strong>and</strong> improvement treatments (precommercial thinning <strong>and</strong> thinning).<br />

Plots affected by fires (612 plots) were excluded from the analysis to avoid confounding effects<br />

between silvicultural treatments <strong>and</strong> fires in those plots.<br />

To quantify afforestation, we calculated the absolute variation of forest area (Δ<strong>Forest</strong>) obtained<br />

from the L<strong>and</strong>-use Map of Catalonia for 1987 <strong>and</strong> 1997 (see methods in Viñas <strong>and</strong> Baulies<br />

1995), subtracting the amount of burnt forest area gathered from the government data of fire<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Gil-Tena et al. 2010. Disentangling recent changes in forest bird ranges in Mediterranean forests (NE Spain)<br />

565<br />

perimeters. Fires were assessed by means of the accumulated amount of burnt forest area<br />

(Burnt) during the first <strong>and</strong> second edition of the CBBA. Furthermore, to determine whether the<br />

initial conditions of forest influence bird range changes, we also computed the initial state of<br />

both basal <strong>and</strong> forest area (G <strong>and</strong> <strong>Forest</strong>, respectively).<br />

2.3 Analysis<br />

We performed Pearson’s correlations between the variation of forest bird species richness <strong>and</strong><br />

the initial forest conditions (<strong>Forest</strong> <strong>and</strong> G), the variation of basal area (ΔG) <strong>and</strong> forest area<br />

(Δ<strong>Forest</strong>), burnt forest (Burnt) <strong>and</strong> the removed basal area in improvement <strong>and</strong> regeneration<br />

treatments (IMP <strong>and</strong> REG Removed G, respectively). Excluding the case of variation of forest<br />

area that in Catalonia is mainly associated with afforestation in formerly cultivated areas, we<br />

computed partial correlations controlling for the effect of initial forest conditions since changes<br />

in forest structure due to forest maturation, fires <strong>and</strong> forest management may be strongly<br />

influenced by them. In the case of partial correlations of the initial forest conditions, for forest<br />

area we considered the influence that may have the initial basal area <strong>and</strong> vice versa. We also<br />

calculated a multivariate general linear model for assessing forest bird species richness variation<br />

as a function of the former independent variables. We used a hierarchical modeling approach<br />

(two steps; see also Gil-Tena et al. (2009, 2010)) to progressively consider initial forest<br />

conditions (Step 1) <strong>and</strong> forest dynamics (Step 2). Significant variables at Step 1 were introduced<br />

in Step 2 <strong>and</strong> their contribution to the model evaluated by means of st<strong>and</strong>ardized partial<br />

regression coefficients (β). In each step, we used a forward stepwise selection process (p-toenter=0.05).<br />

3. Result<br />

The variation of forest bird species richness during the period between Atlases was positive for<br />

much of the study area covered by forests (59% of UTMs covered by forests increased species<br />

richness) whereas the negative values mainly corresponded with poorly forested areas (Figure 1).<br />

Agreeing with this, initial forest conditions were significantly related to the variation of forest<br />

bird species richness (Table 2). In the case of initial forest area (<strong>Forest</strong>), Pearson’s correlation<br />

was significantly different from partial correlation controlling by the effect of initial basal area<br />

(G; p=0.005). This agrees with the positive correlations between variation of species richness<br />

<strong>and</strong> initial basal area (Table 2) <strong>and</strong> with the fact that the difference between Pearson’s <strong>and</strong><br />

partial correlations was lower than for initial forest (p=0.026), indicating that the variation of<br />

species richness was mainly associated with forests with a certain structural development (in<br />

terms of G).<br />

Variation of species richness was also associated with forest dynamics, particularly with<br />

variation of basal area (ΔG). According to the difference between Pearson’s <strong>and</strong> partial<br />

correlations when controlling for the effect of initial basal area (p=0.007), the maturation in<br />

previous forest areas with a developed structure strongly influenced species richness variation.<br />

Afforestation, in terms of increment of forest area, was also positively related with species<br />

richness variation but to a lesser extent than maturation (Table 2). Burnt forest area showed a<br />

non-significant positive correlation with species richness variation that only turned significant<br />

when considering initial forest conditions (Table 2). Regarding forest management influence on<br />

bird species variation, sylvicultural treatments seem not to preclude species richness variation,<br />

which is particularly true for regeneration treatments (Table 2). These results were quite<br />

independent of the initial forest conditions (in all the cases, p>0.05 when comparing Pearson’s<br />

<strong>and</strong> partial correlations).<br />

The results obtained in the multivariate model (R 2 = 0.24; Table 3) confirmed the importance, for<br />

species richness variation, of the maturation produced in forests with a developed structure (in<br />

terms of G). Afforestation also seems to be favoring species richness variation but changes in<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Gil-Tena et al. 2010. Disentangling recent changes in forest bird ranges in Mediterranean forests (NE Spain)<br />

566<br />

forest structure due to management did not appear to have any influence. The weak positive<br />

influence of burnt forest area may indicate that species richness varied despite fires.<br />

4. Discussion<br />

Although the unexplained variation of the models may evidence that other factors not<br />

considered here affected forest bird range changes at the study scale or others, our results<br />

demonstrated that the general positive variation of forest bird species richness (Figure 1) was<br />

significant in forested areas with a developed structure that matured during the 20 last years of<br />

the 20th century. Afforestation was also shown as potentially influence species variation. These<br />

results agree with previous studies at the same scale <strong>and</strong> area (Gil-Tena et al. 2009, 2010) which<br />

showed that forest maturation <strong>and</strong> afforestation, mainly due to l<strong>and</strong> ab<strong>and</strong>onment, have favored<br />

ranges of many forest birds <strong>and</strong> overridden the potentially negative effects of fires. Nevertheless,<br />

we cannot discount different levels of impact of forest fires on forest birds at smaller scales<br />

(


A. Gil-Tena et al. 2010. Disentangling recent changes in forest bird ranges in Mediterranean forests (NE Spain)<br />

567<br />

Gil-Tena, A., Brotons, L. <strong>and</strong> Saura, S., 2009. Mediterranean forest dynamics <strong>and</strong> forest bird<br />

distribution changes in the late 20th century. <strong>Global</strong> <strong>Change</strong> Biology, 15: 474-485.<br />

Gil-Tena, A., Brotons, L. <strong>and</strong> Saura, S., 2010. Effects of forest l<strong>and</strong>scape change <strong>and</strong><br />

management on the range expansion of forest bird species in the Mediterranean region.<br />

<strong>Forest</strong> Ecology <strong>and</strong> Management, 259: 1338-1346.<br />

Gil-Tena, A., Saura, S. <strong>and</strong> Brotons, L., 2007. Effects of forest composition <strong>and</strong> structure on<br />

bird species richness in a Mediterranean context: Implications for forest ecosystem<br />

management. <strong>Forest</strong> Ecology <strong>and</strong> Management, 242: 470-476.<br />

Metzger, M.J., Bunce, R.G.H., Leemans, R. <strong>and</strong> Viner, D., 2008. Projected environmental shifts<br />

under climate change: European trends <strong>and</strong> regional impacts. Environmental<br />

Conservation, 35: 64-75.<br />

Ministerio de Medio Ambiente, 1997-2007. Tercer Inventario <strong>Forest</strong>al Nacional. Dirección<br />

General de Conservación de la Naturaleza, Madrid.<br />

Opdam, P. <strong>and</strong> Wascher, D., 2004. Climate change meets habitat fragmentation: linking<br />

l<strong>and</strong>scape <strong>and</strong> biogeographical scale levels in research <strong>and</strong> conservation. Biological<br />

Conservation, 117: 285-297.<br />

Poyatos, R., Latron, J. <strong>and</strong> Llorens, P., 2003. L<strong>and</strong> use <strong>and</strong> l<strong>and</strong> cover change after agricultural<br />

ab<strong>and</strong>onment: the case of aMediterranean Mountain area (Catalan Pre-Pyrenees).<br />

Mountain Research <strong>and</strong> Development, 23: 362-368.<br />

Sala, O.E., Chapin III, F.S., Armesto, J.J., Berlow, R., Bloomfield, J., Dirzo, R., Huber-<br />

Sanwald, E., Huenneke, L.F., Jackson, R.B., Kinzig, A., Leemans, R., Lodge, D.,<br />

Mooney, H.A., Oesterheld, M., Poff, N.L., Sykes, M.T., Walker, B.H., Walker, M., Wall,<br />

D.H., 2000. <strong>Global</strong> biodiversity scenarios for the year 2100. Science, 287: 1770-1774.<br />

Saura, S. <strong>and</strong> Torné, J., 2009. Conefor Sensinode 2.2: A software package for quantifying the<br />

importance of habitat patches for l<strong>and</strong>scape connectivity. Environmental Modelling<br />

Software, 24: 135-139.<br />

Ukmar, E., Battisti, C., Luiselli, L. <strong>and</strong> Bologna, M.A., 2007. The effects of fire on<br />

communities, guilds <strong>and</strong> species of breeding birds in burnt <strong>and</strong> control pinewoods in<br />

central Italy. Biodiversity <strong>and</strong> Conservation, 16: 3287-3300.<br />

Viñas, O. <strong>and</strong> Baulies, X., 1995. 1:250,000 l<strong>and</strong>-use map of Catalonia (32,000 km2) using<br />

multi-temporal L<strong>and</strong>sat-TM data. International Journal of Remote Sensing, 16: 129-146.<br />

Table 1: Considered forest bird species. According to the CBBA, the trend considering sampling effort is<br />

also reported in brackets. +, - <strong>and</strong> =: exp<strong>and</strong>ing, contracting <strong>and</strong> maintenance trend.<br />

<strong>Forest</strong> bird species (trend)<br />

Accipiter gentilis (-), Aegithalos caudatus (=), Anthus trivialis (=), Caprimulgus europaeus (+),<br />

Circaetus gallicus (+), Corvus corax (=), Corvus corone (=), Dendrocopos major (+), Dendrocopos<br />

minor (+), Dryocopus martius (+), Erithacus rubecula (+), Falco subbuteo (+), Fringilla coelebs (=),<br />

Hieraaetus pennatus (+), Lullula arborea (+), Otus scops (-), Parus ater (+), Parus caeruleus (-), Parus<br />

cristatus(+), Parus palustris (=), Phylloscopus collybita (+), Picus viridis (-), Regulus ignicapilla (+),<br />

Sitta europaea (+), Strix aluco (=), Sylvia atricapilla (+), Sylvia cantillans (+), Troglodytes troglodytes<br />

(-), Turdus philomelos (+), Turdus viscivorus (-)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Gil-Tena et al. 2010. Disentangling recent changes in forest bird ranges in Mediterranean forests (NE Spain)<br />

568<br />

Table 2: Pearson’s <strong>and</strong> partial correlations, when controlling by the effect of the initial forest area (r forest )<br />

<strong>and</strong> basal area (G; r G ), between the change in forest bird species richness <strong>and</strong> forest dynamics, fires <strong>and</strong><br />

management. Δ: <strong>Change</strong>; IMP <strong>and</strong> REG: improvement <strong>and</strong> regeneration treatments, respectively; *, **<br />

<strong>and</strong> ***: p≤0.05, p≤0.01 <strong>and</strong> p≤0.001, respectively.<br />

Variable<br />

Correlation<br />

<strong>Forest</strong> r=0.37***, r G =0.16**<br />

G r=0.41***, r forest =0.25***<br />

∆<strong>Forest</strong><br />

r=0.22***<br />

∆G r=0.44***, r forest =0.31***, r G =0.25***<br />

Burnt r=0.06, r forest =0.11*, r G =0.12*<br />

IMP Removed G r=0.17**, r forest =0.08, r G =0.06<br />

REG Removed G r=0.27***, r forest =0.13*, r G =0.12*<br />

Table 3: Analysis of the factors behind changes in forest bird species richness between the two atlas<br />

periods. The general linear model was conducted in two steps according to a hierarchical process. The<br />

information in the table concerns to that of each variable in their corresponding step of assessment.<br />

β<br />

p<br />

STEP 1: Initial conditions G 0.297


M.B. Horta & E. Keizer 2010. Assessment of human <strong>and</strong> physical factors influencing distribution of vegetation degradation<br />

569<br />

Assessment of human <strong>and</strong> physical factors influencing<br />

spatial distribution of vegetation degradation -<br />

Environmental Protection Area Cachoeira das Andorinhas, Brazil<br />

Marise Barreiros Horta 1* & Edwin Keizer 2<br />

1 Companhia Botânica, Brazil<br />

2<br />

Instituto Nacional de Pesquisas da Amazônia (INPA), Brazil<br />

Abstract<br />

This study examined human <strong>and</strong> physical factors influencing the spatial distribution of<br />

vegetation degradation in a protection area. A map data set was used for the human <strong>and</strong> physical<br />

factors investigation. Those factors comprised: roads network, rural settlements/village/city,<br />

tourist sites, mining sites, agricultural areas, drainage, slope <strong>and</strong> geology. The vegetation<br />

degradation diagnosis was made with utilization of five ecological indicators: cover of invasive<br />

species, understory, canopy, bare soil <strong>and</strong> dead shrub percentage. Regression <strong>and</strong> correlation<br />

analyses were used to investigate the relationship between vegetation degradation <strong>and</strong> factors.<br />

The factor slope presented significantly negatively correlated to vegetation degradation in forest<br />

areas. Distance to tourist sites showed significant negative correlation in the savannah <strong>and</strong> rocky<br />

shrubl<strong>and</strong>s. Those factors can enhance humans <strong>and</strong> livestock accessibility to natural vegetation<br />

areas, which may increase intensity of damaging activities. The information can contribute to<br />

conservation strategies improvements in the protection area.<br />

Keywords: vegetation degradation, change, spatially explicit factors, ecological indicators,<br />

conservation<br />

1. Introduction<br />

Vegetation degradation, unlike deforestation, is not a very obvious phenomenon. The changes<br />

are revealed gradually, sometimes not in terms of decrease of area, but represented by<br />

qualitative losses, for example, through the reduction of species diversity, increase of invasive<br />

species, decrease of the shrub layer, reduction of woody species <strong>and</strong> biomass decline (Hargyono<br />

1993).<br />

The investigations of spatially explicit factors (such as slope, roads) influencing vegetation<br />

degradation processes rely on the fact that those factors can represent an expression to some<br />

underlying structural driving forces, such as demographic <strong>and</strong> political forces (Mather 1990).<br />

Most of the investigation of factors influencing vegetation degradation in the spatial context has<br />

been directed at arid l<strong>and</strong>scapes associated with l<strong>and</strong> degradation, desertification <strong>and</strong> soil<br />

erosion processes (Guerrero-Campo <strong>and</strong> Montserrat-Marti 2000, Kembron 2001). The situation<br />

has generated studies that combine vegetation degradation with other processes, such as, soil<br />

erosion. In general, however, they lack considerations of the quality <strong>and</strong> quantity of vegetation<br />

(Eswaran 2001).<br />

* Corresponding author. Tel.: 55 31 33445037; 55 31 88835937<br />

Email address: marisehorta@companhiabotanica.com.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta & E. Keizer 2010. Assessment of human <strong>and</strong> physical factors influencing distribution of vegetation degradation<br />

570<br />

The present study examined the vegetation degradation phenomenon in a protected area,<br />

characterized by subtropical moderately humid climate, where degradation affects not only<br />

forest, but also other vegetation types. The main objectives of this study were:<br />

1. assess the variations of vegetation degradation;<br />

2. investigate the association between spatial distribution of vegetation degradation <strong>and</strong><br />

human (roads network, rural settlements/village/city, tourist sites, mining sites,<br />

agricultural areas) <strong>and</strong> physical factors (slope, geology, drainage).<br />

2. Methodology<br />

2.1 Study area <strong>and</strong> map data set<br />

The Environmental Protection Area (EPA) Cachoeira das Andorinhas comprises an area of<br />

18.700 ha. It is located in the north of the Ouro Preto Mountain Range, Minas Gerais state, in<br />

the west extreme of the Brazilian Atlantic <strong>Forest</strong> dominion, setting bounds with the Savannah<br />

dominion (Rizzini 1979). The climate is subtropical moderately humid. The mean annual<br />

temperature varies from 19.5ºC to 21.8ºC.<br />

The data sources utilized for the map data set composition comprehended: L<strong>and</strong>sat TM image<br />

30m resolution; topographic map (for roads, rural settlements, villages, <strong>and</strong> city); contour map;<br />

geology map; <strong>and</strong> drainage map. Data collection in the field was carried out for the ground truth<br />

<strong>and</strong> localization of tourist <strong>and</strong> mining sites. The DEM (Digital Elevation Model) was obtained<br />

from a digitized contour map aiming the creation of the slope map. The classification of the<br />

L<strong>and</strong>sat TM image into l<strong>and</strong> cover classes was carried out through a cluster of steps of<br />

supervised classification. The overall classification accuracy obtained was 82%.<br />

2.2. Data collection <strong>and</strong> analysis<br />

Data collection in the field was carried out in a total of 47 sample plots (25 x 25m) using<br />

stratified r<strong>and</strong>om sampling. For the establishment of vegetation degradation levels in the area 5<br />

ecological indicators (I) were selected according to the literature (De Pietri 1992; De Pietri<br />

1995; Hargyono 1993). The classes of vegetation degradation defined comprehended: not<br />

degraded (0), low degraded (1), moderately degraded (2), highly degraded (3) <strong>and</strong> extremely<br />

degraded (4). The direction of the influence of each indicator in the vegetation degradation<br />

process was defined through the use of multivariate analysis (Principal Component Analysis).<br />

Higher proportion of invasive species cover (%), lower estimation of canopy cover (%) <strong>and</strong><br />

lower proportion of understory cover (%) meant higher level of degradation in the forest areas.<br />

The higher degradation condition in the savannah <strong>and</strong> rocky shrubl<strong>and</strong>s was determined<br />

according to the higher proportion of invasive species cover (%), higher proportion of bare soil<br />

(%) <strong>and</strong> higher percentage of dead shrubs. The invasive species considered were: Arundinaria<br />

effusa, Eupatorium sp., Gleichenia sp., Lantana lilacina, Melinis minutiflora, Panicum sp.,<br />

Pteridium aquilinuum, Rhynchospora exaltata, Solanum sp., Vernonia scorpiodes; Poaceae 1;<br />

Poaceae 2.<br />

Principal Component Analysis (PCA) was implemented for the definition of scores that<br />

represented the vegetation degradation variations <strong>and</strong> for checking redundancy in the data. For<br />

the forest areas the human factors distance to agricultural areas <strong>and</strong> distance to tourist sites<br />

presented clustered to village/city/ rural settlements, <strong>and</strong> were removed from the analysis. For<br />

the savannah <strong>and</strong> rocky shrubl<strong>and</strong>s the variable distance to village/city/ rural settlements was<br />

clustered to mining sites <strong>and</strong> removed from the analysis.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta & E. Keizer 2010. Assessment of human <strong>and</strong> physical factors influencing distribution of vegetation degradation<br />

571<br />

The presence <strong>and</strong> absence of damaging activities was verified through the marks <strong>and</strong> signs of<br />

fire, cutting, free grazing, fences, tracks <strong>and</strong> garbage. In order to investigate the relationship<br />

among the variations in vegetation degradation (scores) <strong>and</strong> human <strong>and</strong> physical factors,<br />

regression <strong>and</strong> correlation analysis were performed using SPSS software. The significant factors<br />

comprised those that presented p values significant at α = 0.05, for a one-tailed test.<br />

3. Result<br />

The results of the categorization of the 28 sample plots of the forest areas into vegetation<br />

degradation conditions are presented in the Table 1. From the total of areas sampled, 28% were<br />

classified as extremely degraded (4), 36% as highly degraded (3), <strong>and</strong> 36% were found<br />

moderately degraded (2). The forest intermediate stage (FIS) presented 60% of the sample plots<br />

classified as highly degraded <strong>and</strong> 40% as moderately degraded. The forest advanced stage<br />

(FAS), otherwise, comprehended 40% of sample units classified as highly degraded <strong>and</strong> 60% as<br />

moderately degraded. The scrub areas (S) were all (100%) classified as extremely degraded.<br />

For the savannah (SA) <strong>and</strong> rocky shrubl<strong>and</strong>s formations (RS), in the totality of 19 sample plots,<br />

10 % were found extremely degraded (4), 6% presented highly degraded (3), 37% were<br />

classified as moderately degraded (2), 37% low degraded (1) <strong>and</strong> 10% not degraded (0) (Table<br />

2). The savannah had 9% of the sample plots classified as highly degraded, 46% as moderately<br />

degraded, 36% as low degraded <strong>and</strong> 9% as not degraded. The rocky shrubl<strong>and</strong>s comprehended<br />

25% classified as extremely<br />

The results of the correlation analysis, obtained from linear regression, for investigation of the<br />

relationship among vegetation degradation scores <strong>and</strong> human <strong>and</strong> physical factors, in the forest<br />

areas, are shown in the Table 3. The physical factor slope presented a significant negative<br />

correlation coefficient (R2 = - 0.223; p = 0.011) to vegetation degradation in the forest areas.<br />

The correlation analysis results for the savannah <strong>and</strong> rocky shrubl<strong>and</strong>s formations (Table 4)<br />

showed that the human factor distance to tourist sites was the one that presented a significant<br />

correlation (R2 = - 0.250; p = 0.029) to vegetation degradation.<br />

Activities of cutting <strong>and</strong> grazing were found in higher proportion in the FIS (60% <strong>and</strong> 50%<br />

respectively) <strong>and</strong> FAS (40% <strong>and</strong> 70%), while the presence of fire was evidenced only in the FIS<br />

(10%). Signs of fire occurred in higher proportion in the scrub (100%), savannah (100%) <strong>and</strong><br />

rocky shrubl<strong>and</strong>s (100%). Grazing activities were also testified as important in those areas,<br />

occurring in 88% of the scrub, 90% of the savannah <strong>and</strong> 75% of the rocky shrubl<strong>and</strong>s areas.<br />

Indicators of mining activities, on the other h<strong>and</strong>, were restricted to the rocky shrubl<strong>and</strong>s areas<br />

(25%), due to the presence of the rock substrate, especially quartzite, target of exploitation.<br />

Besides that, garbage signs (cans, plastics, etc) resulted from tourist activities were observed<br />

only in the rocky shrubl<strong>and</strong>s (25%). From the 47 sampled areas 87% showed the presence of<br />

tracks, <strong>and</strong> 81% the absence of fences.<br />

Table 1: Categorization of sample plots of the forest areas according to vegetation degradation indicators<br />

(Ia – Invasive Species Cover; Ib - Understory Cover; Ic – Canopy Cover)<br />

Sample<br />

Vegetation Degradation Vegetation Degradation<br />

Type Ia Ib Ic<br />

Plots<br />

Scores<br />

Classes<br />

1 FIS 2 3 2 -6.816 3<br />

2 FIS 2 2 2 -10.504 2<br />

3 FIS 3 3 2 20.553 3<br />

4 FIS 3 3 2 27.488 3<br />

5 FIS 2 3 3 6.914 3<br />

6 FIS 1 3 2 -21.307 2<br />

7 FIS 3 3 2 29.753 3<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta & E. Keizer 2010. Assessment of human <strong>and</strong> physical factors influencing distribution of vegetation degradation<br />

572<br />

8 FIS 2 1 2 -19.634 2<br />

9 FIS 1 3 2 -19.053 2<br />

10 FIS 2 3 2 0.049 3<br />

11 FAS 3 3 2 20.553 3<br />

12 FAS 3 3 2 32.007 3<br />

13 FAS 1 2 1 -47.215 2<br />

14 FAS 1 1 2 -37.944 2<br />

15 FAS 1 3 1 -30.508 2<br />

16 FAS 1 3 2 -15.294 2<br />

17 FAS 2 3 2 8.456 3<br />

18 FAS 1 3 2 -7.658 2<br />

19 FAS 2 3 2 3.085 3<br />

20 FAS 1 3 2 -3.057 2<br />

21 S 4 0 3 51.892 4<br />

22 S 4 0 3 58.792 4<br />

23 S 4 0 3 58.792 4<br />

24 S 3 0 4 37.400 4<br />

25 S 3 0 4 41.219 4<br />

26 S 4 0 4 64.910 4<br />

27 S 4 0 4 61.092 4<br />

28 S 4 0 4 61.092 4<br />

Table 2: Categorization of sample plots of the savannah <strong>and</strong> rocky shrubl<strong>and</strong>s according to vegetation<br />

degradation indicators (Ia – Invasive Species Cover; Id – Bare Soil Cover; Ie – Dead Shrubs Percentage )<br />

Sample<br />

Vegetation Degradation Vegetation Degradation<br />

Type Ia Id Ie<br />

Plots<br />

Scores<br />

Classes<br />

1 SA 3 2 2 70.965 3<br />

2 SA 3 1 1 62.608 2<br />

3 SA 0 0 3 1.425 1<br />

4 SA 3 0 1 48.332 2<br />

5 SA 2 1 0 30.965 1<br />

6 SA 1 1 2 20.719 2<br />

7 SA 0 2 2 22.127 2<br />

8 SA 2 1 1 34.830 2<br />

9 SA 1 1 1 0.570 1<br />

10 SA 0 1 0 3.640 1<br />

11 SA 0 0 0 0.000 0<br />

12 RS 1 0 1 3.815 1<br />

13 RS 0 0 2 1.083 1<br />

14 RS 0 0 3 1.824 1<br />

15 RS 0 0 0 0.000 0<br />

16 RS 1 2 2 29.471 2<br />

17 RS 1 2 2 36.360 2<br />

18 RS 2 4 2 100.823 4<br />

19 RS 2 4 4 93.542 4<br />

Table 3: Correlation coefficients among human <strong>and</strong> physical factors <strong>and</strong> vegetation degradation<br />

scores in forest areas (bold entry indicate result significant at ∞ = 0.05)<br />

Distance to<br />

village/city/<br />

settlement<br />

Human Factors<br />

Distance to the<br />

roads<br />

Independent Variables<br />

Distance to<br />

mining<br />

sites<br />

Slope %<br />

Physical Factors<br />

Geology<br />

Classes<br />

Distances<br />

to drainage<br />

-0.086 -0.001 -0.021 -0.223 0.072 0.100<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta & E. Keizer 2010. Assessment of human <strong>and</strong> physical factors influencing distribution of vegetation degradation<br />

573<br />

Table 4: Correlation coefficients among human <strong>and</strong> physical factors <strong>and</strong> vegetation degradation<br />

scores in savannah <strong>and</strong> rocky shrubl<strong>and</strong>s (bold entry indicate result significant at ∞ = 0.05)<br />

Independent Variables<br />

Distance to<br />

agricultural<br />

areas<br />

4. Discussion<br />

Human Factors<br />

Distance to the<br />

roads<br />

Distance to<br />

mining<br />

sites<br />

Distance to<br />

tourist sites Slope %<br />

Physical Factors<br />

Geology<br />

Classes<br />

Distances<br />

to drainage<br />

-0.069 -0.137 -0.056 -0.250 -0.181 -0.061 -0.001<br />

The results of categorization of vegetation in degradation classes showed that the majority of<br />

sampled sites were found degraded. The findings have certainly relation with the large amount<br />

of damaging activities signs found in the sampled areas. Different authors refer to the<br />

contribution of activities such as grazing, cutting <strong>and</strong> fire to processes of vegetation degradation<br />

(Kakembo 2001; Hofstad 1997; De Pietri 1995; Kumar <strong>and</strong> Bh<strong>and</strong>ari 1992; De Pietri 1992;<br />

Talbot 1986). Regarding the fencing system Kumar & Bh<strong>and</strong>ari (1992) found inside a<br />

protection site, higher degradation due to free grazing in unfenced areas in relation to the fenced<br />

ones.<br />

The forest areas presented higher levels of degradation while undisturbed <strong>and</strong> low degraded<br />

situations were found only in the savannah <strong>and</strong> rocky shrubl<strong>and</strong>s. The large availability of<br />

resources, especially wood for fuelwood <strong>and</strong> building materials in the forest areas (Michael<br />

Arnold & Dewees 1997), can be a source of major attraction for damaging activities in those<br />

areas.<br />

In the savannah areas of the EPA the trees are short <strong>and</strong> occur in low proportion, sparsely<br />

distributed through the grass layer, <strong>and</strong> consequently they do not have the potential for cutting<br />

<strong>and</strong> charcoal production activities, characterized as the highest important disturbance pressure in<br />

most of the savannah areas in Brazil (Mistry 2000). Disturbances caused by fire, although can<br />

occur in savannah areas in cases of accidental or criminal intensive burning, are not a major<br />

problem when that factor is kept in periodical lower frequencies (Coutinho 1990). Since fire is<br />

an old component of savannah ecosystems the vegetation has developed resistance <strong>and</strong><br />

dependency on this factor (Mistry 2000; Coutinho 1990; Rizzini 1979). Perhaps grazing can<br />

contribute more to degradation processes in the savannah of the EPA, <strong>and</strong> the highest level of<br />

degradation was found in one plot with presence of grazing marks <strong>and</strong> livestock.<br />

The rocky shrubl<strong>and</strong>s are attractive especially for mining activities, but the impacts, although<br />

large, are restricted to the mining location. Similarly, the impacts of tourist activity on the rocky<br />

shrubl<strong>and</strong>s are limited to those areas close to waterfalls <strong>and</strong> cities. Grazing activities, although<br />

evidenced might be limited by the presence of large rock blocks, escarpments <strong>and</strong> deep holes<br />

(Verweij 1995).<br />

The results of correlation analysis for the forest areas showed that slope is a significant physical<br />

factor influencing the vegetation degradation distribution in the EPA Cachoeira das Andorinhas.<br />

The negative association of this factor with forest degradation agrees with Mather (1990). This<br />

author argues that slope is a proximate factor that can increase accessibility of humans to forest<br />

areas. Verweij (1995) argued that cattle also tend to choose relatively easy walking routes to<br />

meet their goals. In the study area, activities of cutting <strong>and</strong> grazing might be hindered by the<br />

slope steepness, with presence of higher degradation levels in the areas with lower slopes.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


M.B. Horta & E. Keizer 2010. Assessment of human <strong>and</strong> physical factors influencing distribution of vegetation degradation<br />

574<br />

In the savannah <strong>and</strong> rocky shrubl<strong>and</strong>s formations the human factor distance to tourist sites<br />

presented a significant negative correlation with vegetation degradation. The tourist visitation is<br />

higher in the southern part of the EPA, where Ouro Preto county <strong>and</strong> the Andorinhas waterfall<br />

are located. In protected areas in Costa Rica <strong>and</strong> Belize, Farrell <strong>and</strong> Marion (2001) also found<br />

that intense tourist visitation can degrade natural resources <strong>and</strong> contribute to vegetation damage<br />

<strong>and</strong> loss. The authors argued that successful ecotourism <strong>and</strong> protected area management needs<br />

effective management of natural areas for visitor enjoyment <strong>and</strong> resource protection.<br />

References<br />

Coutinho, L. M., 1990. Fire in Ecology of the Brazilian Cerrado. In: J. G. Goldammer (Ed.).<br />

Fire in the Tropical Biota - Ecosystem Processes <strong>and</strong> <strong>Global</strong> Challenges (Vol. 84).<br />

Berlin: Springer-Verlag.<br />

De Pietri, D. E., 1992. The search for ecological indicators: is it possible to biomonitor forest<br />

system degradation caused by cattle ranching activities in Argentina Vegetatio : acta<br />

geobotanica, 101(2), 109-121 (114).<br />

De Pietri, D. E., 1995. The spatial configuration of vegetation as an indicator of l<strong>and</strong>scape<br />

degradation due to livestock enterprises in Argentina. Journal of Applied Ecology, 32,<br />

857-865.<br />

Eswaran, H., Lal, R. <strong>and</strong> Reich, P. F., 2000. L<strong>and</strong> degradation: An overview. In: E. Michael<br />

Bridges, R. N. Leslie, T. Compo <strong>and</strong> A. Prueskspong (Eds.). Response to L<strong>and</strong><br />

Degradation. Enfield, New Hampshire: Science Publishers, Inc.<br />

Farrell, T. A. <strong>and</strong> Marion, J. L., 2001. Identifying <strong>and</strong> assessing ecotourism visitor impacts at<br />

eight protected areas in Costa Rica <strong>and</strong> Belize. Environmental Conservation, 28(3), 215-<br />

225.<br />

Guerrero-Campo, J. <strong>and</strong> Montserrat-Martí, G., 2000. Effects of soil erosion on the floristic<br />

composition of plant communities on marl in northeast Spain. Journal of Vegetation<br />

Science, 11, 329-336.<br />

Hargyono., 1993. Occurence <strong>and</strong> prediction of forest degradation : a case study of Upper Konto<br />

watershed, East Java, Indonesia. Unpublished MSc-thesis, ITC - International Institute<br />

for Aerospace Survey <strong>and</strong> Earth Sciences, Enschede.<br />

Hofstad, O., 1997. Degradation processes in Tanzanian woodl<strong>and</strong>s. Forum for Development<br />

Studies, 1, 95-115.<br />

Jongman, R. H., ter Braak, C. J. F. <strong>and</strong> van Tongeren, O. F. R., 1987. Data analysis in<br />

community <strong>and</strong> l<strong>and</strong>scape ecology. Wageningen: Pudoc.<br />

Kakembo, V., 2001. Trends in vegetation degradation in relation to l<strong>and</strong> tenure, rainfall, <strong>and</strong><br />

population changes in Peddle district, Eastern Cape, South Africa. Environmental-<br />

Management, 28(1): 39-46.<br />

Kembron, T., 2001. Natural regeneration of degraded hillslopes in Southern Wello, Ethiopia: a<br />

study based on permanent plots. Applied Geography, 21(3), 275-300.<br />

Mather, A. S., 1990. <strong>Global</strong> forest resources. London: Belhaven Press.<br />

Michael Arnold, J. E. <strong>and</strong> Dewees, P. A., 1997. Farms, Trees & Farmers. Response to<br />

Agricultural Intensification. London: Earthscan Publications LTD.<br />

Mistry, J., 2000. World Savannas - Ecology <strong>and</strong> Human Use. Harlow, Engl<strong>and</strong>: Pearson<br />

Education Limited.<br />

Rizzini, C. T., 1979. Tratado de Fitogeografia do Brasil. Aspectos sociológicos e florísticos.<br />

( Vol. 2). São Paulo: Hucitec - Edusp.<br />

Talbot, L. M., 1986. Demographic factors in resource depletion <strong>and</strong> environmental degradation<br />

in east African rangel<strong>and</strong>. Population <strong>and</strong> developement review, 12(3), 441-451.<br />

Verweij, P.A., 1995. Spatial <strong>and</strong> Temporal Modelling of Vegetation Patterns – Burning <strong>and</strong><br />

grazing in the paramo of Los Nevados National Park, Colombia. Phd-thesis, ITC -<br />

International Institute for Aerospace Survey <strong>and</strong> Earth Sciences, Enschede.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi & M. Akbarzadeh 2010. Role of planted forests <strong>and</strong> trees outside forests in sustainable forest management<br />

575<br />

Role of planted forests <strong>and</strong> trees outside forests in sustainable forest<br />

management in Iran<br />

E. Kouhgardi 1* & M.Akbarzadeh 2<br />

1 Islamic Azad University, Boushehr Branch, Boushehr, Iran.<br />

2 Islamic Azad University, Myianeh Branch, Myianeh, East Azarbaijan, Iran<br />

Abstract<br />

From a forestry point of view, Iran is divided into five vegetation regions <strong>and</strong> during the period<br />

1960-1999, afforestation amounted to 2,221,000 ha total area planted. The annual rate forest<br />

plantations establishment is 63,000 ha, the majority being implemented under government<br />

investment. The state grants substantial support to promote private investment in fast growing<br />

tree species plantations, which amount to 150,000 ha of which, 35% are young st<strong>and</strong>s. The<br />

present evaluation of Trees outside forests in Iran is incomplete for lack of comprehensive data<br />

<strong>and</strong> information. Collaborative efforts between government, municipalities, NGOs <strong>and</strong> citizens’<br />

groups have led to the establishment of a quite dense network of urban <strong>and</strong> peri-urban forests in<br />

Iran, estimated in 1996 to be 530,288 ha. Result show that the current situation of Iran’s natural<br />

resources is a reflection of its past <strong>and</strong> present social, ecological, technological, economic,<br />

political <strong>and</strong> administrative measures. Technical or engineering solutions are not enough; they<br />

need to take into account the needs, priorities <strong>and</strong> aspirations of the rural poor.<br />

Keywords: Plantation, <strong>Forest</strong>, Tree, Urban, Establishment<br />

1. Introduction<br />

Iran covers an area of 1.65 million km2, enclosed within 8,731 km of frontiers, of which 2,700<br />

of coastline boundaries, <strong>and</strong> 6,031 of l<strong>and</strong> borders. Almost 60% of the country is mountainous,<br />

while deserts of the High Central Plateau cover one third of the territory. Environmental<br />

characteristics are Owing to its highly contrasted topography, Iran displays a variety of climates<br />

ranging from hyper arid (centre <strong>and</strong> east regions) to Mediterranean semi-arid <strong>and</strong> sub-humid<br />

(mountain regions) <strong>and</strong> humid (Caspian coastal area, west Azerbaijan <strong>and</strong> southwest Zagros<br />

mountain range). With its mean annual rainfall of 253 mm, Iran is drought-prone; precipitations<br />

being erratic <strong>and</strong> highly variable.<br />

Being endowed with a rich diversity of ecosystems, plant <strong>and</strong> animal species, Iran is one of the<br />

world’s most important gene pools; it counts 8,200 plant species (1,900 endemic), over 500 bird<br />

species <strong>and</strong> 160 species of mammals. Five plant species, 20 mammals, 14 birds, 8 reptiles, 2<br />

amphibians, 7 fish <strong>and</strong> 3 invertebrates are considered either endangered, or threatened <strong>and</strong><br />

vulnerable. Historical evidence indicates that the vast, arid areas of central Iran were once<br />

covered with valuable range <strong>and</strong> forest vegetation. Human activities are believed to have<br />

strongly contributed to desertification. The main l<strong>and</strong> use categories of Iran are the following:<br />

<strong>Forest</strong>s [12.4 million ha - 7.4% of territory]; Rangel<strong>and</strong>s [90 million ha - 55% of the country];<br />

Deserts [34 million ha - 21% of the country]; Cultivated l<strong>and</strong>s [23.6 million ha – 14.4 % of<br />

territory]) exceed by far the forest l<strong>and</strong> area; Urban <strong>and</strong> rural settlements, infrastructures <strong>and</strong><br />

* Corresponding author. Tel.: +98 917 775 9894 - Fax: +98 771 568 3700<br />

Email address: kouhgardi@iaubushehr.ac.ir<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi & M. Akbarzadeh 2010. Role of planted forests <strong>and</strong> trees outside forests in sustainable forest management<br />

576<br />

water bodies (4 million ha - 2.2% of the national territory). The total potential arable l<strong>and</strong><br />

amounts to 37 million ha (17 million ha under irrigation – 20 million ha rain fed).<br />

From view point of surface water <strong>and</strong> groundwater resources it’s divided into 37 basins <strong>and</strong> 174<br />

watersheds, the country is drained by 3,450 permanent <strong>and</strong> seasonal rivers. The Persian Gulf<br />

<strong>and</strong> the Caspian Sea receive the highest flows. In 1996-97 precipitation generated 330 billion<br />

cubic metres (BCM) of surface water, 130 BCM renewable water <strong>and</strong> 126 BCM harvestable<br />

water resources, of which 87.5 BCM were harvested (94 % used by agriculture). About 70<br />

BCM groundwater, were discharged in 1996 by 275,300 semi-deep wells, 100,700 deep wells,<br />

46,700 springs <strong>and</strong> 32,000 qanats.<br />

2. Methodology<br />

2.1 <strong>Forest</strong>s <strong>and</strong> rangel<strong>and</strong>s global estate<br />

As a result of losing ownership <strong>and</strong> usufruct rights, the ex-owners <strong>and</strong> the traditional forest<br />

dwellers/users lost interest <strong>and</strong> sense of responsibility towards sustaining <strong>and</strong> protecting forests<br />

<strong>and</strong> rangel<strong>and</strong>s, used since without restraint to face growing dem<strong>and</strong>s that came with population<br />

growth. <strong>Forest</strong>s occupy 12.4 million ha (7.4 % of country) <strong>and</strong> include 1.9 million ha of<br />

productive commercial forests. The rest amounts to 5.5 million ha (West <strong>and</strong> Zagros),<br />

2.5 million ha (South <strong>and</strong> desert), <strong>and</strong> 2.5 million ha in other regions. Rangel<strong>and</strong>s include l<strong>and</strong>s<br />

covered by natural grassl<strong>and</strong>, shrub-l<strong>and</strong> <strong>and</strong> a combination of both. Iran’s rangel<strong>and</strong>s occupy<br />

90 million ha (54.8% of country area); the condition of 16% of the rangel<strong>and</strong>s is excellent,<br />

whereas 66% are in favorable to fair condition <strong>and</strong> 18% are in poor <strong>and</strong> degraded form.<br />

2.2 Deforestation<br />

The overall deforestation figure for the period 1958–1994 is widely accepted as being equal to<br />

about 5.6 million ha. The presentation that follows gives the rates of deforestation according to<br />

the widely accepted classification of forests in Iran: Caspian broadleaf deciduous forest (1.5<br />

million ha); Arasbaran broadleaf deciduous forest (100,000 ha); Zagros natural forests (1.7<br />

million ha); Irano-Touranian Central <strong>Forest</strong>s (2 million ha); <strong>and</strong> Semi-savannah subtropical<br />

forests (300,000 ha).<br />

2.3 <strong>Change</strong> in vegetation cover<br />

No assessment has been made with regard to the forest annual cover change. However,<br />

considering that the present deforestation is limited, the average annual plantation rate of<br />

63,000 ha should result in a slight positive change in national vegetation cover.<br />

2.3.1 Natural forests<br />

From a forestry point of view, Iran is divided into five vegetation regions as follows: Hyrcanian<br />

broadleaved forests [1.905,000 ha] along the Caspian coast; Arasbaran forests [150,000 ha] of<br />

North Western Iran; Irano-Touranian arid forests [2,895,000 ha] in the Central Plateau Region;<br />

Zagrosian forests [5,050,000 ha]; <strong>and</strong> Persian Gulf <strong>and</strong> Sea of Oman tropical arid forests<br />

[2,400,000 ha].<br />

2.3.2 Planted forests<br />

During the period 1960-1999, afforestation amounted to 2,221,000 ha total area planted (all<br />

categories inclusive). The annual rate forest plantations establishment is 63,000 ha, the majority<br />

being implemented under government investment. Tree species planted are generally limited to<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi & M. Akbarzadeh 2010. Role of planted forests <strong>and</strong> trees outside forests in sustainable forest management<br />

577<br />

indigenous or acclimatized exotic species. To ensure maximum success, most plantations are<br />

irrigated during 2-3 seasons. Water shortages are a major constraint to planting, particularly in<br />

arid zones. Site preparation costs are high, <strong>and</strong> establishment of irrigation facilities very<br />

expensive. The State grants substantial support to promote private investment in fast growing<br />

tree species plantations (poplar), which amount to 150,000 ha of which, 35% are young st<strong>and</strong>s.<br />

2.3.3 Trees outside forests<br />

The present evaluation of trees outside forests in Iran is incomplete for lack of comprehensive<br />

data <strong>and</strong> information. In 2000 it was estimated that orchards accounted for 1,704,000 ha, about<br />

14 % of the total forest area of Iran. Collaborative efforts between government, municipalities,<br />

NGOs <strong>and</strong> citizens’ groups have led to the establishment of a quite dense network of urban <strong>and</strong><br />

peri-urban forests in Iran, estimated in 1996 to be 530,288 ha (mean annual area treated 3,760<br />

ha). Urban <strong>and</strong> peri-urban forestry is gaining momentum in the country <strong>and</strong> many provinces<br />

have developed their own urban forestry establishment program.<br />

2.4 <strong>Forest</strong>, range <strong>and</strong> environmental protection strategies<br />

<strong>Forest</strong>s <strong>and</strong> range: The government is pursuing a strategy of multiple forest utilization <strong>and</strong> is<br />

launching a vigorous national reforestation <strong>and</strong> afforestation program to reclaim degraded<br />

forests <strong>and</strong> rangel<strong>and</strong>s, protect watersheds <strong>and</strong> manage industrial forests on a sustained-yield<br />

basis. It also aims to involve private enterprises to obtain long-term concessions for large forest<br />

areas, with the objective of industrial utilization <strong>and</strong> sustained yield management. In the tree<br />

plantation program, the objective is to move towards more people participation <strong>and</strong> involvement<br />

as several programs are carried out on sub-contract basis with private enterprises.<br />

Environmental protection: The national environmental protection strategy’s goal is to put 10%<br />

of the national l<strong>and</strong> area under protection. At present, 8.2 million ha are being protected by the<br />

Department of Environment, which manages the following categories of protected areas:<br />

National parks (11 sites – 1.3 million ha); Wildlife refuges (25 sites – 1.9 million ha); Protected<br />

areas (47 sites – 5.3 million ha); National nature monuments (5 sites); <strong>and</strong> Biosphere reserves (9<br />

sites – 1.9 million ha).<br />

3. Result<br />

1.1 Causes <strong>and</strong> effects of deforestation <strong>and</strong> degradation<br />

Some of the salient indirect causes to deforestation <strong>and</strong> degradation are:<br />

L<strong>and</strong> <strong>and</strong> water tenure <strong>and</strong> users' rights <strong>and</strong> incentives: These include: Incentives granted to<br />

enhance agricultural production, which have constituted an encouragement to extend the areas<br />

cultivated by converting, with the State’s consent significant forest <strong>and</strong> rangel<strong>and</strong>s to agriculture<br />

l<strong>and</strong>; Indirect incentives granted for forest <strong>and</strong> rangel<strong>and</strong> exploitation, through omitting to tax<br />

products <strong>and</strong> incomes derived from such operations; Promoting excessive water mobilization as<br />

an encouragement to the extension of productive irrigated agriculture, mostly implemented<br />

through forest <strong>and</strong> rangel<strong>and</strong> clearing; <strong>and</strong> L<strong>and</strong> nationalization, which is held responsible for<br />

the breakdown of traditional systems of forest <strong>and</strong> range management, resulting in the<br />

disintegration of the forest <strong>and</strong> range resources.<br />

Poverty triggering factors: These include: Unchecked population growth which inevitably<br />

results in extreme pressure being exerted on the limited natural resource base available to the<br />

country; Poverty, which concerns 40 % of the rural population that attempt to maintain basic<br />

living st<strong>and</strong>ards by increasing livestock numbers on the already overcrowded rangel<strong>and</strong>s; <strong>and</strong><br />

Lack of investments <strong>and</strong> on off-farm job <strong>and</strong> revenue opportunities that compel more people to<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi & M. Akbarzadeh 2010. Role of planted forests <strong>and</strong> trees outside forests in sustainable forest management<br />

578<br />

depend increasingly on marginal l<strong>and</strong>s, at the expense of former forest <strong>and</strong> rangel<strong>and</strong>s for their<br />

crop production.<br />

Response capacity to forest <strong>and</strong> range misuse issues: The response capacity is weak because:<br />

Inability to react on a timely basis to misuse <strong>and</strong> calamity impacts, for lack of timely <strong>and</strong><br />

reliable data <strong>and</strong> information; Top-down approach adopted to trigger community involvement in<br />

the process of natural resources rehabilitation has not set in motion the required sense of<br />

ownership of, <strong>and</strong> responsibility for, the resource that may lead to sustainable success; <strong>and</strong><br />

Some gaps in knowledge related to natural resources, participatory procedures.<br />

Legal/regulatory tools did not incorporate the human dimension <strong>and</strong> consequently failed to<br />

promote the protection <strong>and</strong> sustainable management of the l<strong>and</strong> resources; The form of<br />

conservatism that prevails within the forestry <strong>and</strong> range sector makes it impossible to delegate<br />

fully the responsibility to its traditional users to administer, manage <strong>and</strong> sustain the resource;<br />

<strong>and</strong> Despite commendable efforts, the government’s commitment to sustainable natural<br />

resources management remains insufficient.<br />

The main direct causes to deforestation <strong>and</strong> range degradation are: climatic conditions, which<br />

limit vegetation’s establishment <strong>and</strong> resilience <strong>and</strong> constitute hostile conditions for<br />

rehabilitation; accentuated topography, which triggers acute erosion processes; properties of soil<br />

which often hamper the emergence <strong>and</strong> survival of natural regeneration; natural destructive<br />

calamities, such as floods, wind, temperature <strong>and</strong> drought extremes; <strong>and</strong> pests <strong>and</strong> diseases.<br />

Allocation of forest, woodl<strong>and</strong> <strong>and</strong> rangel<strong>and</strong> for the purpose of agricultural <strong>and</strong> urban<br />

development; Misuse of forest resources through excessive fuel wood, construction, wood<br />

gathering, grazing <strong>and</strong> browsing; Misuse of rangel<strong>and</strong> resources resulting from increased<br />

livestock numbers, well beyond carrying capacity; Inadequate agricultural practices, particularly<br />

abusive utilization of unsuitable farm machinery for subsistence shifting cultivation combined<br />

with “free grazing” on marginal steep sloped l<strong>and</strong>s; Agricultural l<strong>and</strong> ab<strong>and</strong>onment;<br />

Infrastructure construction; <strong>and</strong> Man-made catastrophes such as fires, wars, refugee influxes.<br />

Deforestation <strong>and</strong> degradation result in loss of l<strong>and</strong> productivity, which is translated by decline<br />

in biomass, in species diversity <strong>and</strong> in genetic resource; decline in habitat caused by loss of<br />

vegetative cover, erosion, salinization, water logging, lowering of water tables ; <strong>and</strong> soil erosion<br />

increase.<br />

Natural resources degradation results in poverty expansion. Besides the traditional “underclass”<br />

often identified among forest <strong>and</strong> rangel<strong>and</strong> dwellers, the new population groups affected by<br />

poverty are the rural-to-urban migrants, the l<strong>and</strong>less <strong>and</strong> near l<strong>and</strong>less, the disabled, <strong>and</strong> the<br />

rural female group. The collapse of various production systems, which are not economically<br />

viable anymore, forces more rural population to migrate to cities. Regarding the extent of<br />

deforestation, clearing forest for agriculture, forage production <strong>and</strong> firewood <strong>and</strong> charcoal has<br />

reduced forests by 30 % over the last 40 years.<br />

4. Discussion<br />

4.1 Development choices<br />

Iranian foresters have gained much valuable experience in such fields as s<strong>and</strong> dune fixation,<br />

mangrove regeneration, poplar <strong>and</strong> other fast-growing species plantation, management of nonwood<br />

forest products, development of extensive urban <strong>and</strong> peri-urban forests, development of<br />

intermediary forms of participation to forest <strong>and</strong> range sustainable management ; Despite<br />

imposing resource rehabilitation programs, government has not significantly <strong>and</strong> sustainable<br />

contributed to poverty alleviation among forest <strong>and</strong> rangel<strong>and</strong> dwellers; Following<br />

nationalization of all l<strong>and</strong>s, rapid population growth as well as unsustainable human activities<br />

<strong>and</strong> other natural causes, Iranian forests <strong>and</strong> rangel<strong>and</strong>s have lost very substantial areas in the<br />

last decades.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi & M. Akbarzadeh 2010. Role of planted forests <strong>and</strong> trees outside forests in sustainable forest management<br />

579<br />

Planning <strong>and</strong> decision-making are highly centralized, leaving little space for provincial <strong>and</strong><br />

local initiatives to program <strong>and</strong> project formulation, planning <strong>and</strong> decision-making;<br />

Need to re-assess the country’s training needs for participatory forest <strong>and</strong> range protection,<br />

rehabilitation, management <strong>and</strong> development;<br />

4.2 Natural resources use <strong>and</strong> management<br />

Socio-economic value of forests <strong>and</strong> rangel<strong>and</strong>s is very significant as more than 5 million<br />

people live in forests or in their vicinity, while 450,000 persons live permanently on the<br />

rangel<strong>and</strong>s; <strong>and</strong> Planted forests are established without any preconceived idea of their future<br />

sustainable management. Trees outside forests are not yet well perceived in terms of their actual<br />

or potential contribution to the national economy <strong>and</strong> to the well being of people.<br />

The current situation of Iran’s natural resources is a reflection of its past <strong>and</strong> present social,<br />

ecological, technological, economic, political <strong>and</strong> administrative measures. Technical or<br />

engineering solutions are not enough; they need to take into account the needs, priorities <strong>and</strong><br />

aspirations of the rural poor.<br />

So, it's recommended that: Adopt participatory planning <strong>and</strong> resource management approaches<br />

to sustainable forest resources management, with due regard for biodiversity conservation;<br />

Assessing <strong>and</strong> monitoring ecosystems; Participatory planning <strong>and</strong> management become a<br />

st<strong>and</strong>ard approach to underst<strong>and</strong> the needs <strong>and</strong> aspirations of communities <strong>and</strong> individual<br />

families to contribute in those matters that directly impact upon their sustainable livelihoods;<br />

Poverty alleviation <strong>and</strong> support to local communities: Enhance <strong>and</strong> promote long-term<br />

employment <strong>and</strong> revenue opportunities among forest <strong>and</strong> rangel<strong>and</strong> dwellers by strengthening<br />

stakeholders’ interest <strong>and</strong> investments in sustainable resources management; <strong>and</strong> Development<br />

<strong>and</strong> widespread distribution of alternative domestic energy: Provide alternative domestic energy<br />

sources to cover the entire rural countryside.<br />

Promote more environmentally <strong>and</strong> people friendly approaches to agricultural development <strong>and</strong><br />

expansion, by adopting efficient <strong>and</strong> non-destructive production systems, particularly in<br />

mountain areas, <strong>and</strong> rehabilitating l<strong>and</strong>s that have been exhausted of their productive potential,<br />

to their initial l<strong>and</strong>-use choice; <strong>and</strong> Improve l<strong>and</strong> productivity <strong>and</strong> soil fertility in rehabilitation<br />

of degraded l<strong>and</strong>s, including incorporating trees <strong>and</strong> planted forests in the l<strong>and</strong>scape.<br />

4.3 Enhancing the role of planted forests<br />

Integrated planted forests in a broader l<strong>and</strong>-use context in an attempt to respond to the priority<br />

needs <strong>and</strong> aspirations of people; Maintain or increase the present rate of<br />

afforestation/reforestation; <strong>and</strong> government prepare arid, semi-arid <strong>and</strong> tropical silvicultural <strong>and</strong><br />

management models as well as guidelines for the rehabilitation, silvicultural treatment,<br />

management <strong>and</strong> development of mangroves, fodder tree plantations, <strong>and</strong> Haloxylon persicum<br />

st<strong>and</strong>s developed through plantations <strong>and</strong> seeding.<br />

Grant more support to farmers to maintain <strong>and</strong> exp<strong>and</strong> poplar <strong>and</strong> fast growing species’<br />

plantings for shade, shelter <strong>and</strong> other uses; Promote trees outside forests in private holdings,<br />

particularly in agroforestry where trees support agriculture <strong>and</strong> livelihoods; Develop an adapted<br />

silviculture for the specific needs of urban/peri-urban forests <strong>and</strong> publish silvicultural guidelines<br />

for various species in different ecological contexts; Consider the productive capacity of the<br />

urban <strong>and</strong> peri-urban forests <strong>and</strong> prepare management plans accordingly; Promote the planting<br />

of trees outside forests, mainly fodder trees in sylvo-pastoral systems; <strong>and</strong> Arrange a short<br />

training course on the silvicultural treatments of fodder tree <strong>and</strong> shrub species <strong>and</strong> on sylvopastoral<br />

management of recently rehabilitated wooded rangel<strong>and</strong>s.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi & M. Akbarzadeh 2010. Role of planted forests <strong>and</strong> trees outside forests in sustainable forest management<br />

580<br />

References<br />

Carle, J., Sadio, S., Bekele, M., <strong>and</strong> Rouchiche, S., 2003. Enhancing the Role of Planted <strong>Forest</strong>s<br />

<strong>and</strong> Trees Outside <strong>Forest</strong>s in Low <strong>Forest</strong> Cover Countries, World <strong>Forest</strong>ry Congress,<br />

Qubec, Canada.<br />

Rouchiche, S. <strong>and</strong> Abid, H., 2003. Role of Planted <strong>Forest</strong>s <strong>and</strong> Trees Outside <strong>Forest</strong>s in<br />

Sustainable <strong>Forest</strong> Management: Republic of Tunisia – Country Case Study. Planted<br />

<strong>Forest</strong>s Working Paper FP/27E, FAO, Rome, Italy.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi et al. 2010. Values of mangroves <strong>and</strong> its interaction with marine ecosystem<br />

581<br />

Values of mangroves <strong>and</strong> its interaction with marine ecosystem<br />

E. Kouhgardi 1* , E. Shakerdargah 1 & M. Akbarzadeh 2<br />

1 Islamic Azad University, Boushehr Branch, Boushehr, Iran.<br />

2 Islamic Azad University, Myianeh Branch, East Azarbaijan, Iran<br />

Abstract<br />

Mangrove forests have been traditionally utilized by the local people for a variety of purposes in<br />

Nayb<strong>and</strong> national park. Values of mangroves are recognized as various benefits: Lumber or<br />

similar construction wood; Poles, fuel wood, fishing gear; Raw materials for the wood-based<br />

industry of various nature <strong>and</strong> including board mills, rayon mills, match factories <strong>and</strong> charcoal<br />

products; Non-timber products including tannin to supply raw materials for leather tanning<br />

industries, fishing net processing units, thatching material for roofing <strong>and</strong> raw materials for<br />

indigenous medicine; Edible products including honey <strong>and</strong> wax, game animals, meat <strong>and</strong> fish,<br />

fruits, drinks <strong>and</strong> sugar, natural spawning ground for fish <strong>and</strong> crustaceans, especially for<br />

shrimps <strong>and</strong> prawns; Contribution to mud flat formation <strong>and</strong> control of erosion; Capability to<br />

check inl<strong>and</strong> salinity intrusion; Enhanced capability to combat the impact of cyclone <strong>and</strong> tidal<br />

surge; Enhanced capability to function as a shelter belt during storms <strong>and</strong> cyclones. So in view<br />

point of these various use <strong>and</strong> benefits for human <strong>and</strong> marine ecosystem, conservation of<br />

mangrove forests would be a main strategy in the area.<br />

Keywords: Marine ecosystem, Benefits, Restoration, Yield, Strategy, Impact<br />

1. Introduction<br />

Mangroves are woody trees or shrubs that grow in Intertidal region along tropical <strong>and</strong><br />

subtropical coasts. With favorable geomorphic conditions, mangroves commonly form<br />

extensive tidal forests in moist, humid equatorial climates. Under these conditions, individual<br />

trees may attain heights of 40-45 meters <strong>and</strong> have stem diameters of more than 1 meter. Tidal<br />

mangrove forests under the favorable conditions of humid production climates <strong>and</strong> low to<br />

moderate soil salinity commonly have high rates of primary production <strong>and</strong> growth that are<br />

equivalent to those of the best terrestrial forests. On the other h<strong>and</strong>, under less favorable<br />

conditions, such as high soil salinity or arid climates, rates of primary production <strong>and</strong> growth<br />

are generally somewhat less. There are approximately 60 species of mangroves World-wide.<br />

About 45 of these occur in the South-east Asian/Western Pacific region, commonly referred to<br />

as the New World. The Old World region of Central <strong>and</strong> South America has 45-species, while<br />

there are about 10-12 species of Africa <strong>and</strong> Arabia. Asia <strong>and</strong> the Western Pacific are thus rich in<br />

terms of mangrove flora.<br />

In addition to differences in species richness between major continental regions, their is also a<br />

marked reduction in the number of species with increasing latitude. Thus, while there are about<br />

35 species on the North-eastern tip of Cape York Peninsula, only 4-5 species of mangrove occur<br />

near Brisbane, reducing to a solitary species, Avicennia marina, at the Southernmost limit of<br />

distribution of mangroves at Corner Inlet on the Southern coastline of mainl<strong>and</strong> Australia. There<br />

* Corresponding author. Tel.: +98 917 775 9894 - Fax: +98 771 568 3700<br />

Email address: kouhgardi@iaubushehr.ac.ir<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi et al. 2010. Values of mangroves <strong>and</strong> its interaction with marine ecosystem<br />

582<br />

is a similar reduction in the number of species with increasing latitude in the Northern<br />

hemisphere.<br />

Mangrove forests are important for a number of reasons:<br />

1. Mangrove forests <strong>and</strong> estuaries are the primary nursery area for a number of commercially<br />

important shrimp, crab <strong>and</strong> fish species. They are also important nursery areas for other<br />

species which are they not used commercially, but which for, part of the food chain for<br />

commercial species offshore.<br />

2. Mangrove vegetation stabilizes shorelines <strong>and</strong> the banks of rivers <strong>and</strong> estuaries, providing<br />

them with some protection from tidal bores, ocean currents <strong>and</strong> storm surges.<br />

3. In many countries of South-east Asia <strong>and</strong> the Pacific, mangroves are used commercially<br />

for the production of timber for building, firewood <strong>and</strong> charcoal. Experience has shown that<br />

when these activities are managed appropriately it is possible to derive timber products from<br />

mangrove forests without significant environmental degradation, <strong>and</strong> while maintaining their<br />

value as a nursery <strong>and</strong> source of food for commercial capture fisheries.<br />

2. Methodology<br />

2.1 Mangrove ecosystem<br />

Due to the projected sea level rise, changes are being caused in the mangrove swamps of Sunder<br />

bans. In some areas, mangroves have died out while in other areas mangrove swamps have<br />

become more saline <strong>and</strong> the composition of species of both fauna <strong>and</strong> flora is changing. It is<br />

therefore proposed to study the ecology of mangroves of various salinity levels-low, medium<br />

<strong>and</strong> high. In these areas it is proposed to study the physicochemical conditions, productivity of<br />

benthos, phyto <strong>and</strong> zooplankton, availability of juveniles of prawns <strong>and</strong> fish <strong>and</strong> use them for<br />

growing. In addition, it is also proposed to study the role of different kinds of ecosystems, for<br />

the kind of species they will support for breeding <strong>and</strong> reproduction, as nursery grounds <strong>and</strong> as<br />

habitat for growing adult fish.<br />

2.2 Function <strong>and</strong> uses of mangroves<br />

Mangrove forest ecosystems fulfill a number of important functions <strong>and</strong> provide a wide range of<br />

services at the local <strong>and</strong> national levels (Box). Fishermen, farmers <strong>and</strong> other rural populations<br />

depend on them as a source of wood (e.g. timber, poles, posts, fuel wood, charcoal) <strong>and</strong> nonwood<br />

forest products (food, thatch – especially from nipa palm – fodder, alcohol, sugar,<br />

medicine <strong>and</strong> honey). Mangroves were also often used for the production of tannin suitable for<br />

leather work <strong>and</strong> for the curing <strong>and</strong> dyeing of fishing nets. However, this production has<br />

declined in recent years, mainly because of the introduction of nylon fishing nets <strong>and</strong> the use of<br />

chrome as the predominant agent for curing leather (FAO, 1994).<br />

Mangroves support the conservation of biological diversity by providing habitats, spawning<br />

grounds, nurseries <strong>and</strong> nutrients for a number of animals. These include several endangered<br />

species <strong>and</strong> range from reptiles (e.g. crocodiles, iguanas <strong>and</strong> snakes) <strong>and</strong> amphibians to<br />

mammals (tigers – including the famous Panthera tigris tigris, the Royal Bengal tiger – deer,<br />

otters, manatees <strong>and</strong> dolphins) <strong>and</strong> birds (herons, egrets, pelicans <strong>and</strong> eagles, to cite just a few).<br />

A wide range of commercial <strong>and</strong> non-commercial fish <strong>and</strong> shellfish also depends on these<br />

coastal forests. The role of mangroves in the marine food chain is crucial. According to<br />

Kapetsky (1985), the average yield of fish <strong>and</strong> shellfish in mangrove areas is about 90 kg per<br />

hectare, with maximum yield of up to 225 kg per hectare (FAO, 1994). When mangrove forests<br />

are destroyed, declines in local fish catches often result. Assessments of the links between<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi et al. 2010. Values of mangroves <strong>and</strong> its interaction with marine ecosystem<br />

583<br />

mangrove forests <strong>and</strong> the fishery sector suggested that for every hectare of forest cleared;<br />

nearby coastal fisheries lose some 480 kg of fish per year (MacKinnon <strong>and</strong> MacKinnon, 1986).<br />

Mangrove ecosystems are also used for aquaculture, both as open-water estuarine mariculture<br />

(e.g. oysters <strong>and</strong> mussels) <strong>and</strong> as pond culture (mainly for shrimps).<br />

Because of its high economic return, shrimp farming has been promoted to boost the national<br />

economy <strong>and</strong> alleviate poverty in several countries. This activity is often an answer to the<br />

financial constraints on many farmers <strong>and</strong> local communities <strong>and</strong> represents a source of<br />

employment. However, if unsustainably planned <strong>and</strong> managed, it can lead to uncontrolled<br />

deforestation <strong>and</strong> to pollution of coastal waters, damaged or totally destroyed coastal<br />

ecosystems <strong>and</strong> the loss of the services <strong>and</strong> benefits provided by mangroves. A series of<br />

international principles for responsible shrimp farming have been prepared (FAO/Network of<br />

Aquaculture Centers in Asia-Pacific/UNEP/ World Bank/Worldwide Fund for Nature, 2006;<br />

FAO, 1995a), with the main aim of offering guidance on reducing the sector’s environmental<br />

impact while boosting its contribution to poverty alleviation. The principles were welcomed by<br />

many countries (FAO, 2006b)) <strong>and</strong> will hopefully provide support to the development of more<br />

ecofriendly shrimp production.<br />

The increasing popularity of ecotourism activities also represents a potentially valuable <strong>and</strong><br />

sustainable source of income for many local populations, especially where the forests are easy<br />

accessible.<br />

Mangroves also help protect coral reefs, sea-grass beds <strong>and</strong> shipping lanes by entrapping upl<strong>and</strong><br />

runoff sediments. This is a key function in preventing <strong>and</strong> reducing coastal erosion <strong>and</strong> provides<br />

nearby communities with protection against the effects of wind, waves <strong>and</strong> water currents. In<br />

the aftermath of the 2004 Indian Ocean tsunami, the protective role of mangroves <strong>and</strong> other<br />

coastal forests <strong>and</strong> trees received considerable attention, both in the press <strong>and</strong> in academic<br />

circles. After more than two years, there are still contrasting views on this issue: eyewitnesses<br />

reported that coastal forests had saved villages from the destruction <strong>and</strong> lives, while some<br />

analyses asserted that elevation <strong>and</strong> distance from the coast were more significant determinants<br />

of protection than the forest cover itself.<br />

Even though additional studies are needed to define specific details <strong>and</strong> limits of this protective<br />

function, the numerous studies <strong>and</strong> workshops undertaken on this topic over the past couple of<br />

years have brought to light a number of interesting factors. Experts <strong>and</strong> scientists agree that<br />

thick <strong>and</strong> dense coastal forest belts, if well designed <strong>and</strong> managed, have the potential to act as<br />

bioshields for the protection of people <strong>and</strong> other assets against some tsunamis <strong>and</strong> other coastal<br />

hazards (i.e. coastal erosion, cyclones, wind <strong>and</strong> salt spray). However, generalizations – <strong>and</strong> the<br />

creation of a false sense of protection provided by these bioshields – should be avoided, because<br />

mangroves <strong>and</strong> other coastal forests are not able to provide effective protection against all levels<br />

of hazards <strong>and</strong> may not be effective as shields against tsunamis as severe as the one that<br />

occurred in 2004. A full description of the factors to be taken into account with regard to<br />

enhancing the protective functions of mangroves <strong>and</strong> other coastal forests goes beyond the<br />

scope of this report. Interested readers are referred to FAO (2007) for further information.<br />

2.3 Undervalued resources<br />

Despite the many services <strong>and</strong> benefits provided by mangroves, these coastal forests have often<br />

been undervalued <strong>and</strong> viewed as wastel<strong>and</strong>s <strong>and</strong> unhealthy environments. The high population<br />

pressures frequently present in coastal zones has in some places led to the conversion of<br />

mangrove areas for urban development. In order to increase food security, boost national<br />

economies <strong>and</strong> improve living st<strong>and</strong>ards, many governments encouraged the development of<br />

shrimp <strong>and</strong> fish farming, agriculture, <strong>and</strong> salt <strong>and</strong> rice production in mangrove areas.<br />

Mangroves have also been fragmented <strong>and</strong> degraded through overexploitation for wood forest<br />

products <strong>and</strong> pollution. Indirectly, habitats have been lost because of dam construction on rivers,<br />

which often diverts water <strong>and</strong> modifies the input of sediments, nutrients <strong>and</strong> freshwater. Even<br />

though dense mangrove forests can be important in coastal protection, natural disasters should<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi et al. 2010. Values of mangroves <strong>and</strong> its interaction with marine ecosystem<br />

584<br />

also be listed among the possible causes of degradation: several tropical countries are frequently<br />

hit by cyclones, typhoons <strong>and</strong> strong winds, <strong>and</strong> the trees in the front lines may be damaged<br />

<strong>and</strong>/or uprooted during these catastrophes.<br />

Over the last few years, however, awareness of the importance <strong>and</strong> value of mangrove<br />

ecosystems has been growing, leading to the preparation <strong>and</strong> implementation of new legislation<br />

<strong>and</strong> to better protection <strong>and</strong> management of mangrove resources. In some countries, restoration<br />

or re-expansion of mangrove areas through natural regeneration or active planting has also been<br />

observed. In addition, many governments are increasingly recognizing the importance of<br />

mangroves to fisheries, forestry, coastal protection <strong>and</strong> wildlife. Despite these positive signs,<br />

much still needs to be done to effectively conserve these vital ecosystems.<br />

3. Result<br />

The mangrove l<strong>and</strong>s that, used to be considered as waste l<strong>and</strong> in the past, have recently been<br />

treated as a valuable ecosystem, especially for their unique features. Mangrove forests have<br />

been traditionally utilized by the local people for a variety of purposes. Values of mangroves are<br />

recognized as various benefits. Study developed in the south west of Iran in Boushehr province<br />

<strong>and</strong> recognized that the forest of the mangrove ecosystem is capable to yield the following<br />

direct benefits:<br />

Lumber or similar construction wood; Poles, fuel wood, fishing gear; Raw materials for the<br />

wood-based industry of various nature <strong>and</strong> including board mills, rayon mills, match factories<br />

<strong>and</strong> charcoal products; Non-timber products including tannin (mostly from bark) to supply raw<br />

materials for leather tanning industries, fishing net processing units, thatching material for<br />

roofing <strong>and</strong> raw materials for indigenous medicine; Edible products including honey <strong>and</strong> wax,<br />

game animals, meat <strong>and</strong> fish, fruits, drinks <strong>and</strong> sugar.<br />

The mangrove ecosystem can yield the following indirect benefits:<br />

Natural spawning ground for fish <strong>and</strong> crustaceans, especially for shrimps <strong>and</strong> prawns;<br />

Contribution to mud flat formation <strong>and</strong> control of erosion; Capability to check inl<strong>and</strong> salinity<br />

intrusion; Enhanced capability to combat the impact of cyclone <strong>and</strong> tidal surge; Enhanced<br />

capability to function as a shelter belt during storms <strong>and</strong> cyclones.<br />

4. Discussion<br />

Sustainable management of mangrove<br />

The mangrove ecosystem is a complex one. It is composed of various inter-related elements in<br />

the l<strong>and</strong> sea interface zone which is linked with other natural systems of the coastal region such<br />

as corals, sea grass, coastal fisheries <strong>and</strong> beach vegetation. The mangrove ecosystem consists of<br />

water, muddy soil, trees, shrubs <strong>and</strong> their associated flora, fauna <strong>and</strong> microbes. It is a very<br />

productive ecosystem sustaining various forms of life. Its waters are nursery grounds for fish,<br />

crustacean <strong>and</strong> mollusk <strong>and</strong> also provide habitat for a wide range of aquatic life, while the l<strong>and</strong><br />

supports a rich <strong>and</strong> diverse flora <strong>and</strong> fauna. It also influences the micro climate, prevents coastal<br />

erosion, enhances accretion <strong>and</strong> combats natural calamities such as cyclones <strong>and</strong> tidal bores.<br />

The concept of mangrove management has considerably evolved, as these formations have<br />

become better understood. Instead of simple management of the first st<strong>and</strong>, it is now realized<br />

that the whole ecosystem must be considered. It was also realized that, due to the diversity of<br />

mangrove formations, specific regulations are essential.<br />

For most of the mangrove areas of the world, "fishery" <strong>and</strong> "forestry" are the two conflicting<br />

dem<strong>and</strong>s on mangrove l<strong>and</strong>s. Apportioning of the mangrove l<strong>and</strong> resource to these two major<br />

uses under the concept of sustainable management of the ecosystem needs further research<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Kouhgardi et al. 2010. Values of mangroves <strong>and</strong> its interaction with marine ecosystem<br />

585<br />

though a ratio of 20:80 is suggested for ponds to mangroves, on 25 ha allocations as "woodlot<br />

silvo-fishery" (Choudhury 1996). It is suggested that the mangrove area that may be sacrificed<br />

for fish ponds, can be calculated by using the following formula (Llaurado <strong>and</strong> Lindquist 1982).<br />

The Max. Area that can be brought under Aquaculture = (S-F) / 2S<br />

where:<br />

S= Value of fish yield per hectare<br />

F= Value of forestry per hectare, which must include the linkage values such contribution<br />

to open fishing, near-shore fishing, erosion control, biodiversity value, eco-tourism,<br />

shelterbelt value, etc.<br />

This formula seems to be too simple <strong>and</strong> is based completely on monetary aspects.<br />

Ecological restoration of mangrove habitat is feasible, has been done on a large scale in various<br />

parts of the world, <strong>and</strong> can be done cost-effectively. The simple application of the five steps to<br />

successful mangrove restoration described here would at least ensure an analytical thought<br />

process <strong>and</strong> less use of “gardening” of mangroves as the solution to all mangrove restoration<br />

problems. At appropriate sites with normal or near-normal tidal hydrology <strong>and</strong> with<br />

establishment of mangroves through natural recruitment or planting, restored mangrove systems<br />

can become indistinguishable from nearby natural mangrove systems within a short time. So in<br />

view point of these various use <strong>and</strong> benefits for human <strong>and</strong> marine ecosystem, conservation of<br />

mangrove forests would be a main strategy in the area.<br />

References<br />

Acosta, A., 1997. Use of multi-mesh gillnets <strong>and</strong> trammel nets to estimate fish species<br />

composition in coral reef <strong>and</strong> mangroves in the southwest coast of Puerto Rico. Caribb<br />

Sciences, 33:45–57.<br />

Alongi, D., 2002. Present state <strong>and</strong> future of the world’s mangrove forests. Environmental<br />

Conservation, 29:331–349.<br />

He, B., Fan, H., <strong>and</strong> Mo, Z., 2001. Study on species diversity of fishes in mangrove area of<br />

Yingluo Bay, Guangxi province. Tropical Oceanography, 20:74–79.<br />

Pittman, S., McAlpine, C., <strong>and</strong> Pittman, K., 2004. Linking fish <strong>and</strong> prawns to their environment:<br />

a hierarchical l<strong>and</strong>scape approach. Marin Ecology, 283:233–254.<br />

Sheaves, M., 2005. Nature <strong>and</strong> consequences of biological connectivity in mangrove systems.<br />

Marin Ecology, 302: 293–305.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Schimanski 2010. The importance of Environmental Education to restore <strong>and</strong> preserve natural <strong>and</strong> cultural heritage<br />

586<br />

The importance of Environmental Education to restore <strong>and</strong> preserve<br />

natural <strong>and</strong> cultural heritage: the case of Pirai da Serra – South of<br />

Brazil<br />

Edina Schimanski *<br />

State University of Ponta Grossa, Brazil<br />

Abstract<br />

Pirai da Serra is located in the south of Brazil <strong>and</strong> it represents an impressive natural <strong>and</strong><br />

cultural heritage. The area of Pirai da Serra is characterized by canyons, escarpments, grassl<strong>and</strong><br />

vegetation, araucaria forests, fauna <strong>and</strong> flora diversity, archaeological material (rock art) <strong>and</strong><br />

cultural traditions. In the last decades, despite some efforts to avoid degradation, it is possible to<br />

observe a permanent transformation in the l<strong>and</strong>scape in terms of environmental collapse. Some<br />

examples of local natural dilapidation are a wide use of l<strong>and</strong> to increase agriculture,<br />

development of exotic species of forest, forest burning <strong>and</strong> scarcity of water, among other<br />

environmental impacts. Considering these circumstances, it is urgent the development of<br />

Environmental Education practices which promote the involvement of the local community in<br />

order to prevent a complete devastation. The local community participation (teachers, students,<br />

parents, people from cultural groups, social movements <strong>and</strong> government, among others) is<br />

essential to promote an environmental practice. This kind of practice must be able to<br />

deconstruct non-environmental actions <strong>and</strong> increase the consciousness of people in relation to<br />

restoration <strong>and</strong> preservation of natural <strong>and</strong> cultural heritage of the area.<br />

Keywords: environmental collapse, cultural heritage, environmental education,<br />

1. Introduction<br />

This essay discusses environmental awareness related to the protection of natural <strong>and</strong> cultural<br />

heritage. It argues that environmental education must be understood as a future-orientated<br />

approach to people’s life in relation to preservation of the planet <strong>and</strong> humankind. Environmental<br />

education represents an important tool for the defence of natural <strong>and</strong> ecological reserves such as<br />

Pirai da Serra, located in the South of Brazil. This essay is organized into three parts. The first<br />

part shows some aspects concerning the methodology of the research <strong>and</strong> discusses about the<br />

importance of environmental awareness <strong>and</strong> environmental education. The second part presents<br />

some considerations with regard to the importance of environmental education to preserve <strong>and</strong><br />

restore natural <strong>and</strong> cultural heritage. Finally, the last part brings about some reflections<br />

concerning the idea of critical environmental education <strong>and</strong> the defence of natural <strong>and</strong><br />

cultural heritage.<br />

2. Typifying the methodology of the study <strong>and</strong> introducing some essential concepts<br />

about Environmental Education<br />

Research practice in education, <strong>and</strong> particularly in environmental education, has adopted a<br />

variety of approaches through different theories <strong>and</strong> methodologies. Distinct perspectives have<br />

been used to conceptualise <strong>and</strong> analyse the development of environmental education practice.<br />

* Email address: edinaschi@hormail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Schimanski 2010. The importance of Environmental Education to restore <strong>and</strong> preserve natural <strong>and</strong> cultural heritage<br />

587<br />

This essay is framed on bibliographical research. In synthesis, bibliographical research can be<br />

understood as a scientific mode to gather data from different theoretical backgrounds. This kind<br />

of research is essential since it permits the identification of the status of knowledge. Indeed, it<br />

provides opportunities for new contributions about the subject.<br />

In addition, the observation methodology was used as a process which operated from an open<br />

system of information. This allowed a significant degree of flexibility in the gathering of<br />

general <strong>and</strong> particular impressions from the fieldwork.<br />

In fact, the use of a flexible design represented an important component of the process. Instead<br />

of formal observation, in which the researcher’s intention is merely to observe the situation<br />

without interfering in the event, this study decided to use participant observation. In the case of<br />

this study, participant observation formed a part of a cognitive process in which the researcher<br />

could realise <strong>and</strong> recognise the main aspects of the context.<br />

2.1Environemtal education to restore <strong>and</strong> preserve natural <strong>and</strong> cultural heritage<br />

“If education is the solution, what is the problem” This enigmatic question, posed by David<br />

Orr 2001, raises a challenge to education in contemporary society. There is no doubt that the<br />

environment <strong>and</strong> the associated harmful consequences of human actions represent one of these<br />

challenges for education. As a result, the reality of the environmental world has emerged as a<br />

complex subject which has to be discussed <strong>and</strong> interpreted in the educational arena.<br />

Nevertheless, the environment <strong>and</strong> education are not simple concepts; rather, both incorporate<br />

distinct notions of the world <strong>and</strong> involve complex political <strong>and</strong> social dimensions. Particularly<br />

these concepts are more intricate when connected to the idea of environmental preservation.<br />

Environmental problems are the result of the collapse of economic rationality in contemporary<br />

society <strong>and</strong> that an environmental crisis has arisen from a crisis of civilisation (Leff 2001). This<br />

implies the recognition that there are some limits to capitalist growth <strong>and</strong> a need for the<br />

construction of a new paradigm of sustainable development.<br />

In fact, despite certain international efforts to reduce environmental problems as well as<br />

inequalities <strong>and</strong> the setting out of guidelines for social <strong>and</strong> ecological welfare, many people<br />

around the world are still living in precarious conditions. Contemporary history has shown that<br />

the promise of a different world (greener world) seems to have been overwhelmed by the<br />

dominant structures of modern economies. Instead of offering alternatives to people <strong>and</strong> nature,<br />

such as the implementation of strategies to avoid consumption, modern society has been<br />

encouraging the exploitation of natural resources <strong>and</strong> is endorsing distinctly imperial positions<br />

in a capitalist society. As an example of this is the growing of cultivated areas of exotic forests<br />

(for wood extraction) in opposition to the safeguarding of native forests.<br />

The contradictions this leads to are evident in this study. No doubt, the world has never before<br />

witnessed such a range of technological advances within ‘the global society’. Rapidly <strong>and</strong><br />

distincly these courses of action interfere in the structure of biodiversity as well as in the<br />

natural organization of ecosystems <strong>and</strong> forest l<strong>and</strong>scapes. In addition, many countries have been<br />

suffering the consequences of centralised politics of exclusion from ‘the global world’ in which<br />

large sectors of the population have not been allowed access to basic environmental resources in<br />

society such as drinkable water.<br />

It is within the complex area that lies between rhetoric <strong>and</strong> the real dimensions of environmental<br />

issues that education has played a role in responding to problematic situations from ecological<br />

<strong>and</strong> social crises. In these circumstances, some specialists have used education as a new<br />

discourse to bring about changes of attitudes <strong>and</strong> behaviour in relation to the natural world. In<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Schimanski 2010. The importance of Environmental Education to restore <strong>and</strong> preserve natural <strong>and</strong> cultural heritage<br />

588<br />

this scenario, environmental education is undertaking an important task with regard to<br />

ecological sustainability. For instance, Agenda 21 (UNCED 1992) is particularly clear in this<br />

regard when it supports the notion that education is an essential prerequisite before one can<br />

begin to contribute to these changes to be brought about.<br />

The orientations of education towards the idea of sustainability of the planet <strong>and</strong> raising<br />

consciousness of this matter have become prominent issues for educational principles in the area<br />

of environmental education. These principles, which are based on statements from the Tbilisi<br />

Conference (UNESCO 1977), envisage environmental education as an important component for<br />

the creation of a sustainable environment <strong>and</strong> a fairer society. In global terms, environmental<br />

education which is associated with ideas of sustainability plays a crucial role in dealing with<br />

environmental problems <strong>and</strong> it has been backed up by international educational, political <strong>and</strong><br />

social discourses from governments <strong>and</strong> society.<br />

The main goal of the next topic is to endorse the position that promotion of new strategies for<br />

fair <strong>and</strong> consistent education with an emphasis on a critical perspective is particularly important<br />

in realising the potential of people to promote environmental actions toward natural <strong>and</strong> cultural<br />

heritage.<br />

3. Environmental Education <strong>and</strong> the protection of natural <strong>and</strong> cultural heritage:<br />

the case of Piraí da Serra – South of Brazil<br />

Pirai da Serra (Pirai means Fish River in indigenous name <strong>and</strong> Serra means Peak in<br />

Portuguese) is located in Parana State, South of Brazil. The area of Pirai da Serra represents an<br />

impressive natural <strong>and</strong> cultural heritage. The place is characterized by canyons, escarpments,<br />

grassl<strong>and</strong> vegetation, araucaria forests, fauna <strong>and</strong> flora diversity, archaeological material (rock<br />

art) <strong>and</strong> vast culture. The cultural tradition of the region come from the beginning of Brazilian<br />

colonization <strong>and</strong> it has a wide range of customs which are expressed in people’s habits (e.g.<br />

specific dishes, religious events, among others)<br />

In the last decades, despite some efforts to avoid degradation, it is possible to observe a<br />

permanent transformation in the l<strong>and</strong>scape in terms of environmental collapse. Some examples<br />

of local natural dilapidation are a wide use of l<strong>and</strong> to increase agriculture, development of exotic<br />

species of forest (e.g. pinus plantation) <strong>and</strong> forest burning, among other environmental impacts.<br />

From the middle of last century it is possible to observe a rapid expansion of industrialization<br />

processes nearby the region which, no doubt, affects the local environment. Rapidly, people<br />

from community have been experienced these changes in their lives. Particularly, it is possible<br />

to points out the scarcity of water <strong>and</strong> a progressive deterioration of the native l<strong>and</strong>scape.<br />

Considering these circumstances, it is urgent the development of environmental education<br />

practices which promote the involvement of the local community in order to prevent full<br />

devastation of the place. The local community participation (teachers, students, parents, people<br />

from cultural groups, social movements <strong>and</strong> government, among others) is essential to promote<br />

an environmental practice. This kind of practice must be able to deconstruct non-environmental<br />

actions <strong>and</strong> increase the consciousness of people in relation to restoration <strong>and</strong> preservation of<br />

natural <strong>and</strong> cultural heritage of the area <strong>and</strong> present sustainable alternative practices.<br />

Sustainability should be understood through a critical approach in intellectual <strong>and</strong> practical<br />

terms, as a counter-hegemonic discourse. This means that sustainability must be seen as a<br />

process where people are able to participate in the social <strong>and</strong> political construction of their<br />

environment. In this way, environmental education plays a fundamental role <strong>and</strong> must go<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Schimanski 2010. The importance of Environmental Education to restore <strong>and</strong> preserve natural <strong>and</strong> cultural heritage<br />

589<br />

beyond idealised forms of imposing behaviour <strong>and</strong> attitudes on others; rather it should<br />

incorporate a critical social underst<strong>and</strong>ing of nature <strong>and</strong> environmental problems.<br />

This is what some authors such as Fien 1993 <strong>and</strong> Huckle 1998 regard as ‘education for the<br />

environment’ <strong>and</strong> what Khan 2004 defines as ‘ecopedagogy’. In addition, this is what this essay<br />

endorses as being ‘critical environmental education’ (Schimanski 2005)<br />

It is important to points out that critical environmental education is an essential strategy. In<br />

other words, it may provide the basis for changing behaviour to create new strategies to improve<br />

environmental conditions, leading to a better quality of life <strong>and</strong> to equity. Following this, a<br />

critical interpretation of social <strong>and</strong> ecological concerns is indispensable. This implies breaking<br />

down traditional attitudes, such as the pretence that superiority can be imposed by a dominant<br />

discourse which does not take into account social <strong>and</strong> cultural differences.<br />

In other words, it is necessary to consider the community as a space for praxis. People should<br />

engage with this <strong>and</strong> consequently it is necessary to create new conditions for social<br />

participation at community. To this extent a “new culture of argument” (Myerson & Rydin<br />

1996) linked with critical thinking can be used to usher in a new democratic culture. The<br />

establishment of an emancipatory or critical interest (Habermas 1983) can provide people with a<br />

new underst<strong>and</strong>ing of how to make environmental interpretations of local <strong>and</strong> global changes.<br />

This can be the base on which to create a new “identity <strong>and</strong> ecological democracy” (Huckle<br />

2001). In other words, environmental education is a form of citizenship education, which may<br />

be able to reinforce the ability of people to reflect on <strong>and</strong> act in society. In this sense, some<br />

points are important to reflect about it as follow:<br />

1. It is necessary to formulate new initiatives for support, which encourage critical<br />

thinking in relation to the natural environment <strong>and</strong> cultural heritage;<br />

2. It is necessary to encourage a kind of critical thinking, which can lead to a new<br />

culture of argument <strong>and</strong> a new democratic culture in society concerning environmental<br />

good practices;<br />

3. It is imperative support the growth of creative <strong>and</strong> interdisciplinary critical<br />

approaches for the protection <strong>and</strong> enhancement of natural <strong>and</strong> cultural heritage areas;<br />

4. It is of fundamental importance to clarify that there should be a link between<br />

environmental education <strong>and</strong> citizenship in a political perspective to maximize the<br />

human <strong>and</strong> social potential required in environmental strategies to the defense of nature;<br />

5. Above all, it is essential that community appreciate that there is a connection between<br />

environmental education <strong>and</strong> social justice. This relationship is underpinned by moral<br />

<strong>and</strong> ethical values, can lead to a new way of improving the quality of life for people.<br />

4. Conclusion<br />

Environmental education needs to provide people with the knowledge, underst<strong>and</strong>ing <strong>and</strong><br />

capacity to promote awareness <strong>and</strong> positive attitudes towards environment. In this sense, one of<br />

the main goals of environmental education is focused on the idea of increasing knowledge to<br />

underst<strong>and</strong> ecological processes as well as to stimulate people to get involved in actions which<br />

can help to overcome environmental challenges.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Schimanski 2010. The importance of Environmental Education to restore <strong>and</strong> preserve natural <strong>and</strong> cultural heritage<br />

590<br />

In relation to the defense of natural <strong>and</strong> cultural heritage, environmental education can be an<br />

important strategy to “give voice to people” from community (Armstrong, Moore, Russell <strong>and</strong><br />

Schimanski 2009) This means that environmental education can improve what some authors<br />

call “the sense of place” (see for instance Lesley P. Curthoys & Brent Cuthbertson 2002) in the<br />

community. That is, education plays a fundamental role regarding environmental issues<br />

helping people to improve their sense of being part of the environment.<br />

In the case of Pirai da Serra, the development of environmental education practice should be<br />

able to create new public <strong>and</strong> social spheres in which people feel themselves as subjects <strong>and</strong> not<br />

merely objects in its relation to the nature. In practical terms, this means that environmental<br />

education is the process of becoming enlightened about environmental issues that can give rise<br />

to a more equitable <strong>and</strong> sustainable society by carrying out concrete actions able to produce<br />

good practices in relation to nature.<br />

As a conclusion it is important to asserts that this study regards environmental education as a<br />

process of enlightenment of knowledge through a practice engaged with problem-solving <strong>and</strong><br />

based on participatory actions. In fact, environmental education should go beyond the<br />

development of skills <strong>and</strong> behaviour in issues regarding environmental protection. Protecting<br />

nature is an essential matter but it cannot be undertaken in isolation from an environmental ethic<br />

concerning political, educational <strong>and</strong> social change.<br />

References<br />

Armstrong, F., Moore, M., Russell, O. Schimanski, E. (2009) Action research for creating<br />

inclusive education. IN: Alur, M.& Timmons, V. Inclusive education across cultures.<br />

Crossing Boundaries, Sharing ideas. London: Sage.<br />

Curthoys, L. & Cuthbertson, B.(2002) Listening to the L<strong>and</strong>scape: Interpretive Planning for<br />

Ecological Literacy Canadian Journal of Environmental Education (CJEE), Vol 7, No 2,<br />

Published in cooperation with the Canadian Network for Environmental Education <strong>and</strong><br />

Communication (EECOM) <strong>and</strong> Lakehead University<br />

Fien, J. (1993) Education for the environment. Critical curriculum theorising <strong>and</strong><br />

Environmental Education. Australia: Deakin University<br />

Habermas, J. (1983) Conhecimento e interesse. In: Os pensadores. São Paulo: Abril Cultural<br />

Huckle, J. (1998) Environmental education – between modern capitalism <strong>and</strong> postmodern<br />

socialism. A reply to Lucie Sauvé. www.ec.gc.ca/eco/education/Paper3/Paper3_e.htm<br />

Huckle, J. (2001) Towards ecological citizenship, in LAMBERT, D. & MACHON, P. (2001)<br />

Citizenship through secondary geography .London: Routledge Falmer<br />

Kahn, R. (2004) Towards ecopedagogy: radicalizing environmental education for the<br />

task ahead. http://getvegan.com/ee.htm<br />

Layargues, P. (2000) Solving local environmental problem in Environmental Education:<br />

a Brazilian case study. Environmental Education Research, Vol.6,no.2.London:<br />

Carfax Publishing<br />

Lefff, E. (2001) Educacao ambiental e desenvolvimento sustentavel, in REIGOTA,M (2001)<br />

Verde cotidiano : o meio ambiente em discussao. Rio de Janeiro :DP&<br />

Melo, M. et all. (2007) Projeto de Pesquisa: Diagnóstico ambiental de Piraí da Serra.<br />

Universidade Estadual de Ponta Grossa, Brazil, 14 p.<br />

Myerson, G. & Rydin, I. (1996). The language of environment. A new rhetoric. London:UCL.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Schimanski 2010. The importance of Environmental Education to restore <strong>and</strong> preserve natural <strong>and</strong> cultural heritage<br />

591<br />

Orr, D. (2001) Preface of STERLING, S. (2001). Sustainable Education. Bristol: JW<br />

Arrowsmith<br />

Schimanski, E. (2005) Developing environmental education in Brazilian primary schools<br />

focused on emancipatory actions <strong>and</strong> ecological citizenship: an action research approach.<br />

Thesis submitted to the degree of PhD in Education. University of London.<br />

UNCED (1992) Agenda 21. Earth Summit 92 Rio de Janeiro. London: The Regency Press.<br />

UNESCO (1977) First Intergovernmental Conference on Environmental Education. Final<br />

Report, Tbilissi, USSR. Paris:Unesco<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.A.M. da Silva & E. Shimanki 2010. The certification of forest management <strong>and</strong> its contribution to the rights of workers<br />

592<br />

The certification of forest management <strong>and</strong> its contribution to the<br />

rights of workers<br />

Lenir Aparecida Mainardes da Silva * & Edina Shimanki<br />

Universidade Estadual de Ponta Grossa, Brazil<br />

Abstract<br />

The certification of forest management must assure that the wood used in products that are<br />

extracted from acceptable environmental st<strong>and</strong>ards of forest organization. In addition, forest<br />

certification must have compliance with related national <strong>and</strong> international law. The FSC (<strong>Forest</strong><br />

Stewardship Council) institutes some criteria to guideline conscientious <strong>and</strong> responsible forest<br />

management. These criteria are grounded by environmental, social, cultural <strong>and</strong> economical<br />

aspects which are essentials to promote a sustainable society. In this sense, this study is framed<br />

on some reflections concerning the certification of forest management in Brazil. Particularly it<br />

addresses the need of assuring the rights of workers, <strong>and</strong> compliance with occupational health<br />

<strong>and</strong> safety measures. The main idea is to present some discussions concerning the respect in<br />

relation to rights of workers in connection with community participation in the management<br />

process to control <strong>and</strong> prevent labor injuries.<br />

Keywords: <strong>Forest</strong> Management; labor legislation <strong>and</strong> certification<br />

1. Introduction<br />

According to the Brazilian Association of Planted <strong>Forest</strong>s (2007), Brazil is one of the sectors<br />

that contribute the most to the country's development in the macroeconomic <strong>and</strong> microeconomic<br />

area. However, keeps in its recent history, the daily experience of child workers, workers in<br />

poor working conditions, health <strong>and</strong> semi-slavery, that are disrespectful to the decent work.<br />

According to Silva (2000) is the dramatic situation of workers in this sector, extensive working<br />

hours, low salaries, the non-compliance with laws, the ignorance from the workers about the<br />

risks they are submitted to, <strong>and</strong> absolute lack of control when it comes to working conditions.<br />

Until recently, accidents <strong>and</strong> occupational diseases from these workes in Brazil were only<br />

partially understood by the public opinion, by institutions that have responsibility in the area<br />

<strong>and</strong> by social workers in general. It is in this context that this paper proposes a reflection about<br />

the socio-cultural aspects observed in the process of certification of forest management in Brazil,<br />

related to the current labor legislation (health <strong>and</strong> safety st<strong>and</strong>ards at work), since under those<br />

rules, are the first steps to the garantee of decent working conditions by male <strong>and</strong> female<br />

workers in the forestry sector or outsourced workers, <strong>and</strong> the relationship with neighboring<br />

communities or close to areas of management. The research has its base on literature <strong>and</strong><br />

documents.<br />

* Corresponding author. Tel.: (42) 3220-3387<br />

Email address: lenir@uepg.br<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.A.M. da Silva & E. Shimanki 2010. The certification of forest management <strong>and</strong> its contribution to the rights of workers<br />

593<br />

2. In search of decent work in the Brazilian <strong>Forest</strong> Sector<br />

The concept of forest management has emerged as a form of control of productive forestry<br />

practices, through the recovery in the market for products tha came from responsible forests<br />

management. To the FSC, the certification Principles <strong>and</strong> Criteria to a given <strong>Forest</strong><br />

Management unit, St (P & C FSC) is applicable to all types of forests (tropical, boreal <strong>and</strong><br />

temperate) <strong>and</strong> management types (native or plantation). In this paper, good management means,<br />

as Viana (2002, p.20) which refers to "the best management practices applicable to a particular<br />

forest management unit, considering its characteristics <strong>and</strong> sociocultural constraints,<br />

environmental <strong>and</strong> economic <strong>and</strong> the technical <strong>and</strong> scientific knowledge existing. "<br />

According to the International Labour Organisation, decent work involves the opportunity to<br />

realize a productive work with equitable remuneration, safety at the workplace <strong>and</strong> social<br />

protection for families, better prospects for personal development <strong>and</strong> social integration,<br />

organization <strong>and</strong> participation in decisions affecting their lives. Considering that in the nineties,<br />

Brazil experienced a productive framework for restructuring, with the reduced presence of<br />

workers in society <strong>and</strong> the state with a consequent decrease of social rights.<br />

This problem accompanied by the employment disruption, <strong>and</strong> the precarious working<br />

conditions, such factors contributed to increase the worker´s health injuries. And that was only<br />

in 1988 that Brazil was known as a Democratic State of Law according to its Constitution, <strong>and</strong><br />

that in its Chapter II, about the Social Rights, that rural workers began having the same rights<br />

that the urban workers had.<br />

Considering that the underst<strong>and</strong>ing of society in recent years, is much more improved about<br />

environmental issues, sustainability <strong>and</strong> current resources for environmental monitoring based<br />

on high technologies, have allowed a closer society accompaniment on forest management <strong>and</strong><br />

their environmental impacts.<br />

According to Busch (2008) it is a challenge for entrepreneurs to show the consumers how these<br />

issues have been object of concern. In this sense, the certification process in the forest<br />

management appears as an excellent response to the consumer as well, contributing to garantee<br />

decent working conditions for workers in this sector.<br />

To certify the management, the FSC uses criteria <strong>and</strong> principles which support the responsible<br />

management in compliance with environmental, socio-cultural <strong>and</strong> economic aspects. The<br />

principles to be observed are ten, however, is the principle number 4 (four) entitled: Community<br />

relations <strong>and</strong> workers' rights. Stipulates that forest management should maintain or enhance<br />

long-term social welfare <strong>and</strong> economic development of forest workers <strong>and</strong> local communities.<br />

As the principles are more evident if there is respect for decent work.<br />

The criteria to be observed <strong>and</strong> / or evidenced in this principle:<br />

1.Job opportunities, training <strong>and</strong> other services should be given to the inserted or<br />

communities or the ones that are adjacent to areas of forest management.<br />

2 <strong>Forest</strong> management should achieve or exceed all applicable laws <strong>and</strong> / or regulations<br />

related to health <strong>and</strong> safety of their workers <strong>and</strong> their families.<br />

3. Must be guaranteed the rights of workers to organize <strong>and</strong> voluntarily negotiate with<br />

their employers, as outlined in Conventions 87 <strong>and</strong> 98 of the International Labour<br />

Organisation (ILO).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.A.M. da Silva & E. Shimanki 2010. The certification of forest management <strong>and</strong> its contribution to the rights of workers<br />

594<br />

4. The planning <strong>and</strong> implementation of forest management activities should incorporate<br />

the results of social impact assessments. The consultation process with the population<br />

<strong>and</strong> the groups that are directly affected by the manegement must be maintained.<br />

5. The appropriate mechanisms to resolve complaints <strong>and</strong> provide fair compensation in<br />

case of loss or damage that affects the legal <strong>and</strong> customary rights, the property, the<br />

natural resources or livelihoods of local people should be adopted. The necessary<br />

actions to avoid such losses or damages must be done.<br />

6. The responsible for the forest management must consider initiatives in the social area<br />

field to be included in planning <strong>and</strong> operations of forest management activities. Must be<br />

maintained <strong>and</strong> confirmed the existence of information <strong>and</strong> clear opportunity to<br />

participate from the local community(ies) that is (are) directly affected for forest<br />

management operations, <strong>and</strong> consider their perspectives on the issues that directly affect<br />

their quality of life.<br />

7. There should be mechanisms for dialogue <strong>and</strong> the resolution of complaints between<br />

the worker <strong>and</strong> the responsible for the forest management unit, including the<br />

representation, which is formally recognized by the workers.<br />

8. Workers must pay the minimum consistent with the market average in the region,<br />

according to the productive activity that is performed.<br />

9. Should not be used child labor, in violation of the Law, in the forest management unit.<br />

The work of the young people, that are in the age of apprentices, is only allowed in nonstrenuous<br />

activities considered by the authorities <strong>and</strong> with the guarantee of access to<br />

education.<br />

10. The female labor during pregnancy <strong>and</strong> breastfeeding should be accompanied by<br />

preventive measures of risks <strong>and</strong> dangers that are inherent to the productive activity that<br />

is performed.<br />

11. In the hypothesis of substantial changes in the employment framework of forest<br />

management unit, the preventive actions to minimize the impacts of layoffs on workers<br />

<strong>and</strong> on the local communities must be taken.<br />

12. The adoption of programs or strategies to flexible the work should not result in harm<br />

to the rights lawfully acquired by forestry workers. The person responsible for forest<br />

management unit must undertake sustained efforts to minimize the differences between<br />

workers <strong>and</strong> the contractors themselves <strong>and</strong> avoid the precarious working conditions.<br />

13. The community's access to management <strong>and</strong> non-predatory collection of forest<br />

products derived from wood or not, is allowed <strong>and</strong> regulated in the places where such<br />

access has existed for legal or historical reasons, by formal permission granted by the<br />

responsible from the forest management unit, in compliance with the property rights.<br />

These thirteen criteria, based on indicators, provide an analysis of the reality of the company<br />

being certified. This way, we observe that in Brazil, the companies that are certificated<br />

appropriate themselves to the principles dem<strong>and</strong>s. Particularly in relation to the employees; they<br />

guarantee the proper registration, with wages in accordance with labor agreements; adapt their<br />

structures to provide safe working conditions with a team consisting of doctors <strong>and</strong> labor<br />

engineers, <strong>and</strong> in case of accidents remains the logistic structure suitable for the service.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.A.M. da Silva & E. Shimanki 2010. The certification of forest management <strong>and</strong> its contribution to the rights of workers<br />

595<br />

Regarding the hiring workforce, companies rely on public programs that intermediate the<br />

workforce, prioritizing the local recruitment. In small companies without certification, it is<br />

common the variation of the workforce, given to the low qualification <strong>and</strong> education, that is a<br />

commodious situation for the employers from small businesses related to the recruitment<br />

process, therefore, with the structural surplus of the workforce, many are those who expect a job<br />

oppening, what reflects in the decrease of salaries. In times of unemployment prevailing in the<br />

country in the nineties many workers were subjected to work in any conditions <strong>and</strong> for any<br />

salary.<br />

When necessary the larger companies invest their own resources in training, <strong>and</strong> other smaller,<br />

have resources of public policy work. As Silva (2000) in research conducted in the south region<br />

of the country shows that the limited supply of skilled workforce is one of the intrinsic factors to<br />

this sector, <strong>and</strong> the average in years of schooling from the workers are five years.<br />

A worker in the forestry sector is mostly male, so women in this sector are more in the planting<br />

of seedlings in nurseries<br />

Tem seus direitos a maternidade e a infância assegurada pelas políticas públicas de saúde. They<br />

have their rights to maternity <strong>and</strong> childhood assured by public health policies. As for the<br />

presence of the children´s <strong>and</strong> adolescent´s work in this sector, Brazil since the late eighties, put<br />

on his political force to eradicate child labor <strong>and</strong> confrontation, <strong>and</strong> one of the first activities to<br />

be tackled was the work of children in the coal production in which htere were a high number of<br />

working children.<br />

Dealing with situations of child labor has been a subject of discussion between civil society <strong>and</strong><br />

state, with the definition of several enforcement actions by the government, improving<br />

education policy <strong>and</strong> income transfer programs that are articulated with social protection of the<br />

poor families.<br />

3.Conclusion<br />

According to FSC the certification produces benefits, that are: better prices, increasing of the<br />

productivity, the improving of the image; the guarantee of origin, the market recognition, social<br />

responsibility, contribution to the cause.And, for workers there is an improvement in security<br />

conditions at work, in working conditions. Since the indicators used to have certification have<br />

as a reference the international conventions.<br />

O Fórum Mundial subordinado ao tema ‘Trabalho Digno para uma <strong>Global</strong>ização Justa,<br />

iniciativa conjunta da OIT e da Presidência Portuguesa da União Europeia. Realizado em 2007,<br />

propõe uma meta simples: criar as condições para que mulheres e homens de todo o mundo<br />

para ter acesso a um trabalho digno e produtivo, em condições de liberdade, equidade,<br />

segurança e dignidade. Como imperativo ético e de cidadania, e também como base para um<br />

desenvolvimento sustentável.<br />

To ensure the management sets its principles <strong>and</strong> criteria for certficação on a tripod that<br />

supports the responsible care in compliance with environmental, socio-cultural <strong>and</strong> economic<br />

aspects. Therefore, this tripod is paved in sustainability <strong>and</strong> without decent work there is no<br />

economic or environmental sustainability<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L.A.M. da Silva & E. Shimanki 2010. The certification of forest management <strong>and</strong> its contribution to the rights of workers<br />

596<br />

References<br />

Brazilian Association of Planted <strong>Forest</strong> Producers (ABRAF) Statistical Yearbook of ABRAF:<br />

base year 2006. [Yearbook on the Internet]. Brasília: 2007. [Accessed on 10 Oct. 2007]<br />

available at http://abraflor.org.br/estatística/anuário-ABRAF-2007.pdf<br />

Brazil, the Constitution of the Federative Republic from Brasil.Curitiba: Paraná oficial Press.<br />

Busch, SE, socio 2008.Responsabilidade of suppliers of certified wood-type planting. [PhD<br />

Thesis. Graduate Program in Public Health, University of Paulo.303p.pdf<br />

ILO World Forum on Decent Work for justa.Intervenção Minister of labor <strong>and</strong> social solidarity:<br />

the closing session, November 2, 2007 [accessed on April 16, 2010] Available at:<br />

http://www.ilo.org/public / Portuguese / region / eurpro / lisbon / pdf /<br />

forum_discurso_mtss.pdf<br />

Principles <strong>and</strong> criteria of FSC certification. [Accessed on April 5, 2010] Available at:<br />

http://www.florestavivaamazonas.org.br/2163.php<br />

Silva, L.A.M. 2000. Health risks to the worker for the Timber Industry in the region of Campos<br />

Gerais-Pr. [Dissertation: Program of Postgraduate Studies in Social Work, Catholic<br />

University of São Paulo, 163p.<br />

Viana, V.M, 2002. Brazilian forests <strong>and</strong> the challenges of sustainable development:<br />

management, certification <strong>and</strong> public policy partners. [Associate Professor Thesis: <strong>Forest</strong><br />

Department of <strong>ESA</strong>LQ-USP-.pdf.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Stryamets et al. 2010. Role of non-wood forest products for sustainable development of rural communities<br />

597<br />

Role of non-wood forest products for sustainable development of rural<br />

communities in countries with a transition: Ukraine as a case study<br />

N. Stryamets 1* , M. Elbakidze 2,3 , P. Angelstam 2 & R. Axelsson 2<br />

1 Ukrainian National <strong>Forest</strong>ry University, Ukraine<br />

2 Swedish University of Agricultural Sciences, Sweden<br />

3 Ivan Franko National University, Ukraine<br />

Abstract<br />

Sustainable use of non-wood forest products (NWFPs) is a component of sustainable forest<br />

management. NWFPs provide important use <strong>and</strong> non-use values to stakeholders in rural<br />

l<strong>and</strong>scapes. The aim of this study was to analyse the policies with relevance for NWFP in order<br />

to define the potential contribution of these forest resources to rural development in countries in<br />

transition from socialistic planned to market economy, like Ukraine. We analysed national <strong>and</strong><br />

international policy documents, national <strong>and</strong> regional management regulations concerning the<br />

use of NWFPs <strong>and</strong> reviewed relevant literature. We concluded that there were a need to<br />

investigate the role of NWFP in countries with different economic <strong>and</strong> social-cultural conditions<br />

<strong>and</strong> systems of governance of natural resources.<br />

Keywords: non-wood forest products, sustainable forest management, countries in transition,<br />

Ukraine<br />

1. Introduction<br />

<strong>Global</strong>ly forests provide wood resources <strong>and</strong> a large variety of non-wood goods as a resource<br />

base for livelihoods <strong>and</strong> development. Non-wood forest products (NWFPs) are defined as goods<br />

of biological origin other than wood, derived from forests, other wooded l<strong>and</strong> <strong>and</strong> trees outside<br />

forests (Ch<strong>and</strong>rasekharan 1995). Use of NWFPs has a long history as an important component<br />

of the livelihoods of people living in <strong>and</strong> in the vicinity of forest <strong>and</strong> woodl<strong>and</strong> l<strong>and</strong>scapes.<br />

NWFPs were, in fact, one of the first sources of food, medicine, fibre, <strong>and</strong> other substances that<br />

have sustained local communities since the first human settlements (Ryabchuk 1996, Bradley<br />

<strong>and</strong> Bennett 2002). Estimates indicate that 80% of the population in developing countries in the<br />

world use NWFPs to meet some of their nutritional needs (Janse <strong>and</strong> Ottitsch 2005). In addition<br />

to food forests also provide herbal medicine which is vital to poor people that are not a part of a<br />

national healthcare system (Ch<strong>and</strong>rasekharan 1995, Bradley <strong>and</strong> Bennett 2002; Ryabchuk 1996;<br />

Malyk 2006).<br />

NWFPs have attracted considerable interest as an important component of sustainable forest<br />

management (SFM) policies (Elbakidze et al 2007, Elbakidze <strong>and</strong> Angelstam 2007, MCPFE<br />

1993, Siry et al 2005), <strong>and</strong> in different implementation initiatives <strong>and</strong> projects towards<br />

sustainable management of forest l<strong>and</strong>scapes in many countries (Varma et al 2000). There have<br />

been many efforts to complement industrial timber production with an aim to increase the<br />

multiple values of forests for users, owners <strong>and</strong> local communities that depend on them.<br />

* Corresponding author. Tel.: +380979603016<br />

Email address: Natalie.Stryamets@smsk.slu.se<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Stryamets et al. 2010. Role of non-wood forest products for sustainable development of rural communities<br />

598<br />

However, the role of NWFP in the livelihoods of local communities in different contexts has not<br />

been studied in detail. In particular there is a lack of comparative studies which show the<br />

contribution of NWFPs in rural development in countries with different economic <strong>and</strong> socialcultural<br />

conditions as well as systems of nature resource governance. Such knowledge is needed<br />

for policy-makers <strong>and</strong> organisations involved in rural <strong>and</strong> economic development, which put<br />

their efforts to promote new modes of forest use replacing or disrupting existing traditional<br />

livelihood strategies (Hyde <strong>and</strong> Köhlin 2000).<br />

Following the break-up of the Soviet Union, Ukraine re-appeared as an independent new state in<br />

1991. To promote sustainability on national as well as regional <strong>and</strong> local levels Ukraine has<br />

joined the process of developing sustainable forest management principles. In Ukraine these are<br />

oriented towards sustainable yield forestry, maintenance of forest biodiversity <strong>and</strong> socio-cultural<br />

values (<strong>Forest</strong> Code of Ukraine 2006). The strategic objectives of the national forest policy are<br />

related to those enumerated in international agreements of sustainable development (UNCED<br />

1992), sustainable use <strong>and</strong> protection of forests in Europe (MCPFE 1995, MCPFE 2003,<br />

MCPFE 2007). Ukraine has signed the 17 resolutions of the Ministerial Conference on<br />

Protection of <strong>Forest</strong>s in Europe.<br />

The end of Soviet central planning caused a decline of jobs in industry including forestry <strong>and</strong><br />

agriculture (Nordberg, 2007). In many forest <strong>and</strong> woodl<strong>and</strong> regions of Ukraine local people<br />

therefore have had to go back to traditional l<strong>and</strong> use practises due to difficult economic<br />

conditions (Elbakidze <strong>and</strong> Angelstam, 2007). Different NWFPs have thus become a part of the<br />

social fabric <strong>and</strong> livelihood (Bihun, 2005) in many rural areas. There is now often conflict<br />

between increase of harvested timber <strong>and</strong> the emerging sustained yield forestry <strong>and</strong> vital<br />

interests of local people in the sustainable production of NWFP.<br />

The aim of this study was to analyse the policies with relevance for NWFP in order to define the<br />

potential contribution of these forest resources to rural development in countries with different<br />

economic <strong>and</strong> social-cultural conditions <strong>and</strong> systems of governance <strong>and</strong> government of natural<br />

resources. We analysed national <strong>and</strong> international policy documents concerning sustainable<br />

forest management, national <strong>and</strong> regional management regulations related to the use of NWFPs<br />

<strong>and</strong> reviewed relevant literature.<br />

As the next steps we will interview local forest stakeholders in villages located in our case study<br />

areas in Ukraine, study the reasons, places <strong>and</strong> methods of NWFPs collection, amount of<br />

harvested NWFPs by different groups of forest stakeholders, existing practices of NWFP<br />

utilization, including traditional one.<br />

2. Methodology<br />

First we analysed national forest policy documents, management regulations concerning use of<br />

forest resources in Ukraine. A comprehensive literature review was done about NWFPs,<br />

management of forest resources <strong>and</strong> rural development.<br />

3. Result<br />

Sustainable use of non-wood forest products as a component of sustainable forest<br />

management policies<br />

Policies at global, international, EU <strong>and</strong> national levels clearly pronounce the importance of<br />

NWFP. Indeed, NWFP as a relevant attribute to rural development <strong>and</strong> natural resource<br />

conservation have increased globally (Janse <strong>and</strong> Ottitsch, 2005, Ch<strong>and</strong>rasekharan 1995, Ticktin<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Stryamets et al. 2010. Role of non-wood forest products for sustainable development of rural communities<br />

599<br />

2004, Elbakidze <strong>and</strong> Angelstam 2009). At the United Nations Conference on Environment <strong>and</strong><br />

Development (UNCED) in 1992 in Rio de Janeiro it was declared that the promotion of <strong>and</strong> use<br />

of non-wood forest products is an important part of sustainable development (UNCED, 1992).<br />

SFM is supported by different processes <strong>and</strong> organizations, taking into account the specific<br />

forest conditions (The Montréal Process 2007, McDonalda <strong>and</strong> Lane 2004, Rametsteiner <strong>and</strong><br />

Mayer 2004) The Montreal Process (MP) developed SFM principles for the temperate <strong>and</strong><br />

boreal forests of non-European countries (The Montréal Process 2007).<br />

Sustainable forest management (SFM), an important part of sustainable development as a<br />

societal process at the Pan-European level, is described as “the stewardship <strong>and</strong> use of forests<br />

<strong>and</strong> forest l<strong>and</strong>s in a way, <strong>and</strong> at a rate, that maintains their biodiversity, productivity,<br />

regeneration capacity, vitality <strong>and</strong> their potential to fulfil, now <strong>and</strong> in the future, relevant<br />

ecological, economic <strong>and</strong> social functions, at local, national, <strong>and</strong> global levels, <strong>and</strong> that does not<br />

cause damage to other ecosystems” in the Helsinki Resolution which is the European level<br />

process to develop the SFM concept (MCPFE 1993).<br />

Criteria <strong>and</strong> indicators for SFM in European counties were developed by the Ministerial<br />

Conferences on Protection of <strong>Forest</strong> in Europe (MCPFE 1993, MCPFE 1998a, MCPFE 1998b,<br />

McDonalda <strong>and</strong> Lane 2004). The European criteria <strong>and</strong> indicators provide guidelines for SFM<br />

at the national <strong>and</strong> sub-national levels, <strong>and</strong> is an attempt to operationalise <strong>and</strong> complement the<br />

existing definition of SFM (Lazdinis, 2000).<br />

Management of NWFPs is a component of SFM according to international policy documents,<br />

such as, for example, MCPFE resolutions. It is declared in the Helsinki resolution (MCPFE<br />

1993) that the interest of <strong>and</strong> dem<strong>and</strong> for non-wood forest products has been increasing <strong>and</strong><br />

encouraged as a part of sustainable management of forests (MCPFE 1993). It also promotes the<br />

cooperation of the forestry sector in developed countries with countries with economic in<br />

transition (MCPFE 1993).<br />

According to the Resolution L2 (MCPFE 1998a), criteria 3 is to maintain <strong>and</strong> encourage<br />

productive functions of forests, which include both wood <strong>and</strong> non-wood products. The<br />

descriptive indicators of the criteria 3 require the development of management plans for NWFP<br />

(MCPFE 1998a). At the 4th Ministerial conference in Vienna the criteria <strong>and</strong> indicators were<br />

improved with the aim to increase benefits of rural livelihoods from forests (MCPFE 2003,<br />

Rametsteiner <strong>and</strong> Mayer 2004, Wang 2004) <strong>and</strong> included values <strong>and</strong> quantity of non-wood<br />

goods from forests <strong>and</strong> other wooded l<strong>and</strong>s (Rametsteiner <strong>and</strong> Mayer 2004). Vienna Resolution<br />

2 highlighted the importance of the promoting the use of wood <strong>and</strong> NWFPs (MCPFE 2003).<br />

Analysis of national forest policy documents <strong>and</strong> management regulations<br />

Ukraine has recently joined the process of developing national SFM principles, which have<br />

been adopted into the national legislation <strong>and</strong> forest programs (<strong>Forest</strong>ry code of Ukraine 2006,<br />

Lisy Ukrainy, 2009). The main trend of the forest legislation development has been to provide a<br />

balance between the conservation of forest ecosystems, <strong>and</strong> sustainable multi-purpose use of<br />

forests (Angelstam et al. 2009). The outcomes of forest sector reformation, which began in 1991,<br />

were new <strong>Forest</strong> Code adaptation in 2006 <strong>and</strong> elaboration of the State Program “<strong>Forest</strong>s of<br />

Ukraine during 2010-2015” in 2009. In Ukraine forest management is conducted according to<br />

the <strong>Forest</strong> Code <strong>and</strong> within the framework defined by the State Programme “<strong>Forest</strong> of Ukraine<br />

in 2010-2015”.<br />

The <strong>Forest</strong>ry Code of Ukraine (2006) consists of 110 articles, of which four include information<br />

about NWFPs <strong>and</strong> their management, these are called secondary forest products. There is,<br />

however, neither an explanation of what kinds of non-wood products that are included into this<br />

category, nor how they should be managed (<strong>Forest</strong>ry Code of Ukraine 2006). The direct use of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Stryamets et al. 2010. Role of non-wood forest products for sustainable development of rural communities<br />

600<br />

NWFP includes harvesting of hay, grazing, picking fruits, nuts, mushrooms, berries <strong>and</strong> medical<br />

herb. The collection of NWFP for private needs in state <strong>and</strong> community forests is free to<br />

everyone. Private forests are owned by private persons or companies, <strong>and</strong> may occupy up to 5<br />

ha. In private forest local people have to secure a permit for harvesting of NWFPs from the<br />

owner (<strong>Forest</strong>ry Code of Ukraine, 2006). Collection of NWFPs with the aim to sell them is<br />

called “special use of NWFPs”. Commercial collection of NWFPs by a private person or a<br />

company requires a special permit <strong>and</strong> the collector has to pay a fee to the state or the forest<br />

owner.<br />

The State program “<strong>Forest</strong>s of Ukraine during 2010-2015” is based on the MCPFE criteria <strong>and</strong><br />

indicators, <strong>and</strong> defines the guidelines for forest management towards SFM. Collection of<br />

secondary forest products (NWFPs) in managed forests should be done without harming forest<br />

ecosystems (Cabinet Ministriv of Ukraine 1996). Medical herbs <strong>and</strong> mushrooms which are<br />

listed in the Red Data Book of Ukraine (Cabinet Ministriv of Ukraine 1996, Red Data Book of<br />

Ukraine, 1996) are not allowed to harvest, not even parts of the plants or mushrooms. There is a<br />

list of species that are endangered <strong>and</strong> should be collected under strict control, this include a<br />

special ticket or permit that all collectors has to buy. Harvesting of wild berries is allowed if the<br />

ground covers of berries plant are more than 10%, the cover of medical herbs are more than 5%<br />

of the ground cover in the forest (Cabinet Ministriv of Ukraine, 1996). The requirements<br />

concerning harvesting of medical herbs also include regulations about the parts of the herbs<br />

which could be collected, for example, roots should be harvested less than 10%. For leaves <strong>and</strong><br />

st<strong>and</strong>s of the herb less than 40% of the biological productivity in the forest could be harvested.<br />

The forestry enterprises are obliged to protect forest wood <strong>and</strong> non-wood resources from illegal<br />

or harmful consumption by people.<br />

In protected forests there are many restrictions regarding harvesting of NWFPs. There are<br />

certain limitations on harvesting of NWFPs in other types of protected areas. For example, in a<br />

national nature park the use of NWFPs is prohibited in the management zone of strict nature<br />

protection, however, it is allowed in the management zones where tourist faculties are located<br />

<strong>and</strong> where local people conduct their l<strong>and</strong> use activities. Collection is always prohibited in strict<br />

protected reserves (Law of Ukraine on Nature Protected Areas in Ukraine, 1992).<br />

4. Discussion<br />

The economic importance of the forest <strong>and</strong> forestry wood <strong>and</strong> non-wood products is significant,<br />

especially in rural areas in the north <strong>and</strong> west of Ukraine (Nordberg, 2007). Commercialisation<br />

of NWFPs has a potential to contribute to the livelihoods of rural people in Ukraine. Valueadded<br />

processing of NWFPs could in addition contribute to household income.<br />

Thus, sustainable management of NWFP is potentially important to support rural development<br />

(MCPFE, 1998). NWFP <strong>and</strong> their value-added processing have attracted considerable interest as<br />

a component of different development projects in recent years due to their potential to support<br />

rural livelihoods (Angelstam et al. 2009, Angelstam <strong>and</strong> Elbakidze 2009). At the same time, in<br />

some European regions, NWFPs <strong>and</strong> services provide more revenue than wood sales (MCPFE<br />

2007, Arnold <strong>and</strong> Pe´rez 2001). However, there are many challenges to balance production of<br />

wood as is economically the most important product of forests <strong>and</strong> the increasing dem<strong>and</strong> for<br />

NWFPs from European forests (MCPFE, 2007).<br />

To conclude, there is a need to investigate the role of NWFP in countries with different<br />

economic <strong>and</strong> social-cultural conditions <strong>and</strong> systems of governance of natural resources. The<br />

next step in our studies of NWFP will be to interview local forest stakeholders in villages<br />

located in case studies in Ukraine <strong>and</strong> Sweden. Key topics that will be studied include the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Stryamets et al. 2010. Role of non-wood forest products for sustainable development of rural communities<br />

601<br />

reasons, places/habitats <strong>and</strong> methods of NWFPs collection, amount of harvested NWFPs by<br />

different groups of forest stakeholders, existing practices of NWFP utilization, including<br />

traditional one.<br />

References<br />

Angelstam, P., Elbakidze, M., 2009. Traditional knowledge for sustainable management of<br />

forest l<strong>and</strong>scapes in Europe’s East <strong>and</strong> West. In: Soloviy, I.P., Keeton, W.S. (eds.).<br />

Ecological Economics <strong>and</strong> Sustainable <strong>Forest</strong> Management: Developing a transdisciplinary<br />

approach for the Carpathian Mountains. Ukrainian National <strong>Forest</strong>ry<br />

University Press, Lviv, Ukraine. pp. 151-162.<br />

Angelstam, P., Elbakidze, M., Axelsson; R., Törnblom, J., 2009. Sustainable <strong>Forest</strong><br />

Management from policy to practice, <strong>and</strong> back again: l<strong>and</strong>scapes as laboratories for<br />

knowledge production <strong>and</strong> learning in the Carpathian Mountains. In: Soloviy, I.P.,<br />

Keeton, W.S. (eds.). Ecological Economics <strong>and</strong> Sustainable <strong>Forest</strong> Management:<br />

Developing a trans-disciplinary approach for the Carpathian Mountains. Ukrainian<br />

National <strong>Forest</strong>ry University Press, Lviv, Ukraine. pp. 389-414.<br />

Arnold, M.J.E., Pe´rez M. R., 2001. ANALYSIS. Can non-timber forest products match tropical<br />

forest conservation <strong>and</strong> development objectives Ecological Economics 39, 437–447.<br />

Bihun, Y., 2005. Principles of Sustainable <strong>Forest</strong> Management in the Framework of Regional<br />

Economic Development. Vistnyk Lvivs’kogo Unviversytetu. Seria Geografichna. 32, pp.<br />

19–32.<br />

Bradley C. Bennett., 2002. <strong>Forest</strong> products <strong>and</strong> traditional peoples: Economic, biological, <strong>and</strong><br />

cultural considerations. Natural Resources Forum 26, 293–301<br />

Cabinet Ministriv of Ukraine, 1996. Regulation on Second-Quality Timber Harvest <strong>and</strong> Non-<br />

Timber Products <strong>and</strong> Services in Ukrainian <strong>Forest</strong>s. Resolution, Cabinet of Ministers,<br />

Ukraine, 23.04.1996 № 449<br />

Ch<strong>and</strong>rasekharan C., 1995 Terminology, definition <strong>and</strong> classification of forest products other<br />

than wood. In: Report. International Expert Consultation on Non-Wood forest products.<br />

Schmincke K.-H. Yogyakarta. Rome, Italy: Food <strong>and</strong> Agriculture organization of the UN<br />

(FAO), <strong>Forest</strong> products division, 1995. 342-380pp.<br />

Elbakidze, M., Angelstam, P., 2007. Implementing sustainable forest management in Ukraine’s<br />

Carpathian Mountains: The role of traditional village systems. <strong>Forest</strong> Ecology <strong>and</strong><br />

Management 249: 28–38.<br />

Elbakidze, M., Angelstam, P., 2009. Role of traditional village systems for sustainable forest<br />

l<strong>and</strong>scapes: a case study in the Ukrainian Carpathian Mountains. In: Soloviy, I.P., Keeton, W.S.<br />

(eds.). Ecological Economics <strong>and</strong> Sustainable <strong>Forest</strong> Management: Developing a transdisciplinary<br />

approach for the Carpathian Mountains. Ukrainian National <strong>Forest</strong>ry University<br />

Press, Lviv, Ukraine. pp. 301-316.<br />

Elbakidze, M., Angelstam, P., Axelsson, R.,2007. Sustainable forest management as an<br />

approach to regional development in the Russian Federation: state <strong>and</strong> trends in<br />

Kovdozersky Model <strong>Forest</strong> in the Barents region. Sc<strong>and</strong>inavian Journal of <strong>Forest</strong><br />

Research 22:568-581.<br />

<strong>Forest</strong>ry Code of Ukraine, 2006. 08.02.2006 .(In Ukrainian)<br />

Gensiruk, S.A., 1992. <strong>Forest</strong>s of Ukraine, Lisy Ykrainy. Naukova dumka, Kiev, 408 pp.(In<br />

Ukrainian)<br />

Hyde, W.F., Köhlin, G.,2000. Social forestry reconsidered. Silva Fennica 34(3), 285–314.<br />

Janse, G., Ottitsch, A., 2005. Factors influencing the role of Non-Wood <strong>Forest</strong> Products <strong>and</strong><br />

Services, <strong>Forest</strong> Policy <strong>and</strong> Economics 7, 309– 319<br />

Law of Ukraine on Nature Protected Areas in Ukraine, 1992 .16.06.92, № 34 .(In Ukrainian)<br />

Lazdinis, M., 2000. Sustainable forest development. Intoductory research essey. Departament<br />

of conservation biology, SLU, Uppsala. 41.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


N. Stryamets et al. 2010. Role of non-wood forest products for sustainable development of rural communities<br />

602<br />

Lisy Ukrainy, 2009 “On State Programme <strong>Forest</strong>s of Ukraine 2010-2015”, 16.09.09 № 977. (In<br />

Ukrainian)<br />

Malyk, L., 2006. Modern consisting <strong>and</strong> problems of the use of non-wood forest resources of<br />

western region of Ukraine. Lviv, Naukovyu visnyk 16.5. (in Ukrainian).<br />

McDonalda, G., Lane M., 2004. Converging global indicators for sustainable forest<br />

management. <strong>Forest</strong> Policy <strong>and</strong> Economics 6, 63–70.<br />

MCPFE, 1993. Resolution H1General Guidelines for the Sustainable Management of <strong>Forest</strong>s in<br />

Europe. Second Ministerial Conference on the Protection of <strong>Forest</strong>s in Europe 16-17 June<br />

1993, Helsinki, Finl<strong>and</strong>. 5p.<br />

MCPFE, 1995. Pan-European Criteria <strong>and</strong> Indicators for Sustainable <strong>Forest</strong> Management.<br />

Annex 1 of the Resolution 1, Lisbon, Vienna Liaison Unit.<br />

MCPFE, 1998a. Annex 1 of the Resolution L2 Pan-European Criteria <strong>and</strong> Indicators for<br />

Sustainable <strong>Forest</strong> Management. Third Ministerial Conference on the Protection of<br />

<strong>Forest</strong>s in Europe. 2-4 June 1998, Lisbon/Portugal 14p.<br />

MCPFE, 1998b. Resolution L1 People, <strong>Forest</strong>s <strong>and</strong> <strong>Forest</strong>ry –Enhancement of Socio-Economic<br />

Aspects of Sustainable <strong>Forest</strong> Management. Third Ministerial Conference on the<br />

Protection of <strong>Forest</strong>s in Europe. 2-4 June 1998, Lisbon/Portugal 4p.<br />

MCPFE, 2003. State of Europe’s <strong>Forest</strong>; Fourth Ministerial Conference on the Protection of<br />

<strong>Forest</strong>s in Europe, 28-30 April 2003.- Liaison Unit, Vienna<br />

MCPFE, 2007. Fifth Ministerial Conference on the Protection of <strong>Forest</strong>s in Europe. Conference<br />

Proceedings, 5–7 November 2007, Warsaw, Pol<strong>and</strong>, 272p<br />

Nordberg, M., 2007. Ukraine reforms in forestry 1990-2000. <strong>Forest</strong> Policy <strong>and</strong> Economics 9:<br />

713-729.<br />

Rametsteiner, E.,Mayer, P.,2004. Sustainable forest management <strong>and</strong> Pan-European forest<br />

policy. – Ecological Bullitien. 51: 51–57.<br />

Red Data Book of Ukraine, 1996. Roslynnuj svit. Kiev, 1996, 605. .(In Ukrainian)<br />

Ryabchuk, V.,1996. Non-wood forest resources. Lviv. 312. (in Ukrainian).<br />

Siry, J., W. Cubbage, F., Ahmed, M. R., 2005. Sustainable forest management: global trends<br />

<strong>and</strong> opportunities <strong>Forest</strong> Policy <strong>and</strong> Economics 7 551– 561<br />

The Montréal Process, 2007. Criteria <strong>and</strong> indicators for the conservation <strong>and</strong> sustainable<br />

management of temperate <strong>and</strong> boreal forests. Third<br />

Edition.http://www.rinya.maff.go.jp/mpci/<br />

Ticktin, T., 2004. The ecological implications of harvesting non-timber forest products, Journal<br />

of Applied Ecology 2004 41, 11–21<br />

United Nations Conference on Environment <strong>and</strong> Development (UNCED). 1992. Report of the<br />

United Nations Conference on environment <strong>and</strong> development, Rio de Janeiro, 3-14 June<br />

1992<br />

Varma, V., Ferguson, I., Wild, I., 2000. Decision support system for the sustainable forest<br />

management. <strong>Forest</strong> Ecology <strong>and</strong> Management 128, 49-55.<br />

Wang, S., 2004. One hundred faces of sustainable forest management. <strong>Forest</strong> Policy <strong>and</strong><br />

Economics 6, 205– 213.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Volkova 2010. The regional analysis of forest management risks<br />

603<br />

The regional analysis of forest management risks<br />

(by the example of Russian northern areas)<br />

E. Volkova<br />

Institute of Monitoring for Climatic <strong>and</strong> Ecological Systems SB RAS, Tomsk, Russia<br />

Abstract<br />

The northern taiga areas of Russia are rich of coniferous <strong>and</strong> softwood forests, but a number of<br />

adverse natural-climatic <strong>and</strong> economic factors makes proper forest management difficult.<br />

Additional efforts <strong>and</strong> measures are required to mitigate the impact of the harsh climatic<br />

conditions on the forestry activities. The analysis of natural hazards <strong>and</strong> risks in forest<br />

management, which would also take into account related social <strong>and</strong> economic consequences, is<br />

a prerequisite condition for sustainable usage <strong>and</strong> preservation of forest potential in the northern<br />

territories of Russia. In order to assess <strong>and</strong> estimate the risks in forest management, there was<br />

conducted an integrated analysis based on weighted summation of such quantitative indices as<br />

the amount of forest resources, ecological potential of woods, the degree of natural-climatic<br />

hazards <strong>and</strong> other factors. Based on the calculation of the forest resource balance, the<br />

quantitative estimation of valuable woods loss was made.<br />

Keywords: forest management risks, natural-climatic hazards, economic consequences.<br />

1. Introduction<br />

The efficiency analysis of forest management practices usually includes the assessment of risks<br />

caused by environmental <strong>and</strong> climatic conditions of the territory. The risk level also largely<br />

depends on economic <strong>and</strong> geographical situation in the region: natural resource potential,<br />

remoteness from woodworking centers, development of roads <strong>and</strong> transportation network,<br />

availability of qualified manpower. Environmental conditions are a major factor determining<br />

how fast or slow the territories are settled <strong>and</strong> cultivated, while social <strong>and</strong> economic situation<br />

mainly influences the scale, forms <strong>and</strong> methods of using natural resources <strong>and</strong> economic<br />

opportunities based on them.<br />

The risks in forest management can be defined as the probability of full or partial destruction of<br />

the region’s forest resource base resulting in considerable economic damage <strong>and</strong> caused both by<br />

the action of natural processes <strong>and</strong> the influence of anthropogenic factors. Research in this<br />

sphere is especially topical for those regions where full-scale economic activity <strong>and</strong> sustainable<br />

forest exploitation are restrained by adverse environmental <strong>and</strong> climatic conditions, despite of<br />

the considerable natural resource potential.<br />

In this respect, the Tomsk region (Tomsk oblast, as an administrative unit) presents a classic<br />

example of a northern Russian territory – with intensive usage of forest resources <strong>and</strong> a high<br />

forest resource potential. The region possesses considerable forest resources – the forest fund<br />

l<strong>and</strong>s occupy 90.5% of the territory; of them the forest covered l<strong>and</strong>s make up 67.3%. Total<br />

amount of timber reserves of the territory is 2820.88 mln. m 3 . Over 70% of the forest fund is<br />

formed by commercial forest, about 30% – by protected forest, of which approximately 6%<br />

belong to specially protected natural territories. Approximately a half of the commercial forest<br />

consists of coniferous woods, most valuable tree species here being pine-tree, fur-tree, silver fir<br />

<strong>and</strong> cedar (<strong>Forest</strong> Plan 2008).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Volkova 2010. The regional analysis of forest management risks<br />

604<br />

Despite such a considerable forest vegetation potential, the share of forestry products in the total<br />

regional product structure accounts only for 6%. In many respects this is due to economic<br />

reasons – general recession in the development of forest industry, distant location of logging<br />

sites, deterioration of the major forest fund <strong>and</strong> other reasons (State of environment 2009).<br />

In addition to economic reasons, severe environmental <strong>and</strong> climatic conditions play there own<br />

restraining role: low winter temperatures, great weather variability, large <strong>and</strong> sudden changes of<br />

daily <strong>and</strong> annual temperatures, severe hydrological situation. All these factors aggravate the<br />

risks of forest exploitation activities in the Tomsk region territory; creation of forestry-based<br />

infrastructure requires involving additional resources in order to minimize the consequences of<br />

adverse weather conditions <strong>and</strong> better adapt to them. Thus, the analysis of natural hazards <strong>and</strong><br />

risks influencing the forest exploitation practices, which would take into account social <strong>and</strong><br />

economic consequences connected with them, is an important <strong>and</strong> essential prerequisite<br />

condition for sustainable usage <strong>and</strong> preservation of the forest resources of the territory.<br />

2. Methodology<br />

This paper presents the analysis of risks in the sphere of forest management, considering both<br />

environmental <strong>and</strong> economic parameters <strong>and</strong> based on the complex analysis of environmental<br />

<strong>and</strong> climatic hazards, as well as resource <strong>and</strong> ecological potential. For the analysis of the study<br />

area the comparative-geographical method was employed. To determine the risks connected<br />

with adverse environmental conditions, we used the methods of qualitative factor analysis <strong>and</strong><br />

ecologo-economic balance of the territory. Also, a considerable amount of reference statistical<br />

<strong>and</strong> cartographic material was used in the research.<br />

Climatic <strong>and</strong> hydrological conditions bearing risks for the forest exploitation in the Tomsk<br />

region can be divided into two classes: the risks connected with plantations loss <strong>and</strong> the risks<br />

connected with forestry works management. The second group of risks associated with<br />

geographical location <strong>and</strong> natural properties of the territory can be divided into hydrological <strong>and</strong><br />

climatic – these risks are mostly indirect, but their socio-economic consequences are no less<br />

important.<br />

To the date, the first group of risks, those leading to the plantations loss, has been studied to<br />

more detail. The greatest risk in the forest exploitation is connected with forest fires, which are<br />

primarily of anthropogenic origin. Also, forest diseases <strong>and</strong> injurious insects have a large<br />

impact. The area of plantings lost for these reasons varies considerably by years <strong>and</strong> different<br />

forestry sites. The loss amount mostly depends on weather conditions of the current year <strong>and</strong>, in<br />

case of harmful insects, also of previous years. Besides, 15% of total risks are connected with<br />

hurricane winds <strong>and</strong> 13% – with forest fires caused by thunderstorms. At this, the main damage<br />

falls onto economic forests of natural origin, where the impact of all adverse factors is revealed<br />

most vividly (Nevidimova <strong>and</strong> Melnik <strong>and</strong> Volkova 2009).<br />

For the analysis of ecological <strong>and</strong> forest resource potential different types of plantings were<br />

chosen, based on the statistical data for the period of 1991-2006 <strong>and</strong> the data received from 21<br />

weather forecast stations located in different climatic areas of the territory.<br />

The assessment of ecological potential included such stratum parameters of dominant species as<br />

productivity class, average volume density <strong>and</strong> average annual growth rate. The tree species<br />

types <strong>and</strong> plantations proportion in the forested area have also been taken into account in the<br />

calculation. Ecological potential can be defined as a set of useful properties of planting<br />

communities, needed to perform their basic environmental, l<strong>and</strong>scape-protecting, l<strong>and</strong>scapestabilizing<br />

<strong>and</strong> economico-ecological functions.<br />

To practically assess the impact of each of these factors on the ecological potential, their<br />

heterogeneous quantitative parameters were converted into scores. Ecological potential (EP)<br />

was calculated by the formula:<br />

n<br />

EP = 2/3 ∑ S<br />

i<br />

( Bi<br />

+ Pi<br />

+ Zi<br />

) +1/3 S<br />

j<br />

( B<br />

j<br />

+ Pj<br />

+ Z<br />

j<br />

)<br />

i=<br />

1<br />

k<br />

∑<br />

j = 1<br />

(1),<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Volkova 2010. The regional analysis of forest management risks<br />

605<br />

where B - average productivity class; P - average volume density, scores; Z - current growth<br />

rate, scores; S - percentage of each species in the forested area. The first component<br />

characterizes the ecological potential of coniferous species, the second one – that of softwood<br />

species. The ratio of deciduous to coniferous woods is one to three, showing a characteristic<br />

share of each species in the Tomsk region. The territorial distribution of the estimated<br />

ecological potential for the Tomsk region is shown in the figure below.<br />

The forest resource potential is understood as a set of useful properties of forest communities<br />

allowing them to provide natural raw materials, food <strong>and</strong> feed resources, to perform industrial<br />

<strong>and</strong> energy supply functions for the economy. When assessing the forest resource potential, the<br />

following indices were considered: average productivity class, average growing stock, average<br />

age <strong>and</strong> species composition of the forest st<strong>and</strong>.<br />

By analogy with the assessment of ecological potential, the forest resource potential (RP) was<br />

calculated based on the growth rate tables <strong>and</strong> using the following formula:<br />

n<br />

∑<br />

RP = 2/3 S<br />

i<br />

( Bi<br />

+ M<br />

i<br />

+ Ai<br />

) +1/3 ∑ S<br />

j<br />

( B<br />

j<br />

+ M<br />

j<br />

+ A<br />

j<br />

)<br />

i=<br />

1<br />

k<br />

j=<br />

1<br />

where B - average productivity class, scores; M - indicator of growing stock, scores; A - average<br />

age, scores; S - percentage of each tree species in the forested area. The first component<br />

characterizes the forest resource potential of coniferous species, <strong>and</strong> the second one – that of<br />

softwood species.<br />

3. Result<br />

Ecological potential <strong>and</strong> resource potential of forests were assessed for all major tree species<br />

present in the Tomsk region. The table 1 shows the assessment of ecological potential <strong>and</strong> forest<br />

resource potential for different forestry sites, based on the example of a pine tree, as one of the<br />

most common tree species in the Tomsk region. Based on the tables of growth rates <strong>and</strong> other<br />

quantitative characteristics, all average stratum parameters used in the analysis were assigned a<br />

score of 1 to 5 (see Table 2).<br />

Integrated risk analysis in forest management largely depends on the resource <strong>and</strong> ecological<br />

potential of forests <strong>and</strong> the impact of climatic factors most dangerous to forests – forest fires,<br />

winds <strong>and</strong> thunderstorms, injurious insects, forest diseases.<br />

The risks of forest fires, winds <strong>and</strong> thunderstorms, pests <strong>and</strong> forest diseases were assessed as a<br />

functional dependence between the level of each type of risks <strong>and</strong> the level of resource <strong>and</strong><br />

ecological potentials of the forests.<br />

Based on the conducted research <strong>and</strong> approbation of the described approach, there was<br />

performed territorial differentiation <strong>and</strong> quantification of the Tomsk region by the risk degree in<br />

forest management (see Fig. 1).<br />

4. Discussion<br />

The research has shown that both the nature <strong>and</strong> the degree of risks in forest management are<br />

determined by climatic <strong>and</strong> forest growing conditions in the Tomsk region. Negative<br />

environmental factors have a large impact, directly affecting the development of forest industry<br />

in the region. However, geographical distribution of forest exploitation activities within the<br />

Tomsk region itself does not largely depend on the climatic conditions, being mainly<br />

determined by the resource potential of forests <strong>and</strong> transport accessibility of the area. Thus, the<br />

main logging activities in the territory of Tomsk region are conducted in the areas with the<br />

highest degree of risks <strong>and</strong> adverse factors in forest exploitation, being drawn by a high timber<br />

resource potential in these areas.<br />

(2),<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Volkova 2010. The regional analysis of forest management risks<br />

606<br />

In addition to the general strategy for sustainable usage of the region’s natural resources based<br />

on the analysis of environmental risks, we have proposed a system of limitations <strong>and</strong> regulations<br />

for forest management practices. To increase the adaptive capacity of forest resources, the<br />

impact of natural risks should be mitigated by taking certain measures. For northern territories,<br />

the following key measures for better adaptation to the major risk factors are recommended:<br />

reduction of deforestation, restoration of cultivated peatl<strong>and</strong>s, tree species improvement with the<br />

purpose to increase the biomass productivity <strong>and</strong> ecological functions of forests, creation of<br />

mineralized strips, periodic tree cuttings for better forest quality <strong>and</strong> thinning, preventive<br />

controlled fires, cleaning the forest from rubbish, creation of prophylactic barriers against the<br />

fire, building roads for fire fighting purposes, extermination of stem pests, etc.<br />

The described above approach can be used in other regions as well, taking into account<br />

specific local environmental conditions.<br />

References<br />

<strong>Forest</strong> Plan of Tomsk oblast, 2008. Tomsk: Department of development <strong>and</strong> business<br />

undertakings: 12-53.<br />

State of environment of Tomsk oblast in 2008, 2009. Tomsk: University Press: 124-205.<br />

Nevidimova O.G., Melnik M.A., Volkova E.S., 2009. Analysis of natural-climatic hazardS oN<br />

Tomsk oblast territory for estimate of the nature management risks. Ecology of the<br />

urbanized territories, 2: 71-77.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Volkova 2010. The regional analysis of forest management risks<br />

607<br />

Table 1: The assessment of ecological potential <strong>and</strong> forest resource potential of a pine tree for different<br />

forestry sites<br />

The forestry sites<br />

Age<br />

Age , score<br />

Productivity class<br />

Productivity class ,<br />

score<br />

An average volume<br />

density<br />

An average volume<br />

density , score<br />

Growing stock<br />

Growing stock , score<br />

Current growth rate<br />

Current growth rate,<br />

score<br />

Percentage of a pine<br />

tree in the forested<br />

The assessment of<br />

ecological potential<br />

The assessment<br />

resource potential of<br />

forests<br />

Aleks<strong>and</strong>rovskoe 138 4 4 2 0,53 3 221 4 1,2 2 0,39 2,76 3,95<br />

Asinovskoe 85 3 3,8 3 0,68 5 146 2 2,1 3 0,21 2,36 1,72<br />

Bakcharskoe 112 5 5,7 1 0,55 3 95 1 0,9 1 0,26 1,29 1,81<br />

Verhneketskoe 126 4 4,3 2 0,61 5 133 2 1,4 2 0,45 4,05 3,60<br />

Kargasokskoe 122 4 4,9 2 0,56 3 126 2 1,2 2 0,37 2,62 3,00<br />

Vasuganskoe 133 4 5,4 1 0,55 3 114 2 0,9 1 0,26 1,28 1,79<br />

Kolpashevskoe 112 5 5,1 1 0,56 3 103 2 1,1 2 0,32 1,94 2,58<br />

Kriivosheinskoe 107 5 4,9 2 0,61 5 120 2 1,4 2 0,23 2,11 2,11<br />

Molchanovskoe 110 5 4,7 2 0,58 3 116 2 1,5 2 0,35 2,43 3,12<br />

Kedrovskoe 118 5 5,4 1 0,54 3 105 2 0,9 1 0,21 1,06 1,70<br />

Parabelskoe 110 5 4,7 2 0,55 3 117 2 1,2 2 0,35 2,44 3,14<br />

Pervomaiskoe 84 3 3,3 3 0,66 5 169 3 2,6 3 0,09 0,96 0,79<br />

Teguldetskoe 120 5 4,7 2 0,6 5 126 2 1,3 2 0,07 0,65 0,65<br />

Chainskoe 123 4 5,8 1 0,49 3 86 1 0,7 1 0,13 0,66 0,79<br />

Shegarskoe 105 5 4,6 2 0,58 3 112 2 1,5 2 0,14 0,96 1,23<br />

Zuryanskoe 80 2 1,8 5 0,72 5 231 4 3,5 5 0,04 0,58 0,43<br />

Timiryazevskoe 84 3 2,4 4 0,74 5 219 4 2,9 4 0,34 4,44 3,76<br />

Tomskoe 77 2 1,4 5 0,73 5 248 4 3,7 5 0,04 0,54 0,39<br />

Table 2: The scale of average stratum parameters in a scores based on progress of growth of a pine tree<br />

Productivity<br />

class Score Age Score<br />

An average<br />

volume<br />

density<br />

Score<br />

Indicator of<br />

growing<br />

stock<br />

Score<br />

Current<br />

growth rate<br />

1-2 5 101-121 5 0,6-0,79 5 more 250 5 more 3,3 5<br />

2-3 4 121-141 4 0,8-1 4 200-249 4 2,5 -3,2 4<br />

3-4 3 81-101 3 0,4-0,59 3 150-199 3 2-2,4 3<br />

4-5 2 61-81 2 0,2-0,39 2 100-149 2 1-1,9 2<br />

more 5 1 41-61 1 0,1-0,19 1 less 100 1 less 1 1<br />

Score<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Volkova 2010. The regional analysis of forest management risks<br />

608<br />

Figure 1: The risks in forest management in the Tomsk region<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Section 8<br />

Urban forestry in changing regions


I. Löfström et al. 2010. Biodiversity <strong>and</strong> recreational values in urban participatory forest planning in Finl<strong>and</strong><br />

610<br />

Biodiversity <strong>and</strong> recreational values in urban participatory forest<br />

planning in Finl<strong>and</strong><br />

Irja Löfström 1* , Mikko Kurttila 2 , Leena Hamberg 1 & Laura Nieminen 1<br />

1<br />

Finnish <strong>Forest</strong> Research Institute, P.O. Box 18, FIN-01301, Vantaa, Finl<strong>and</strong><br />

2<br />

Finnish <strong>Forest</strong> Research Institute, P.O. Box 68, FIN-80101, Joensuu, Finl<strong>and</strong><br />

Abstract<br />

In Finl<strong>and</strong> most of the urban green areas are forests with indigenous understorey vegetation.<br />

Urban forests are often located near residential areas, <strong>and</strong> they are actively used for outdoor<br />

recreation. Municipalities own most of these forests in Finl<strong>and</strong>. The aim of this study was to<br />

gain overall picture <strong>and</strong> to discover the main development needs of planning <strong>and</strong> managing<br />

municipality owned urban forest.<br />

Most of the municipalities have inventoried the biodiversity of their forests, <strong>and</strong> used lighter<br />

management methods than for commercial forests. Participatory planning has been commonly<br />

used for taking forest users’ opinions into account <strong>and</strong> for integrating multiple values. Multifunctional<br />

forest planning, multi-criteria decision analysis <strong>and</strong> advanced decision-support tools<br />

are still rather new approaches in the field of urban forestry. However, these modern methods<br />

could be useful for integrating multiple values into planning process <strong>and</strong> thus to improve the<br />

quality of urban forest planning.<br />

Keywords: municipally owned urban forest, participatory planning, biodiversity, recreation<br />

1. Introduction<br />

In Finl<strong>and</strong> about 80 % of urban green areas are forests. According to Finnish law, citizens are<br />

allowed to use forests for recreation freely, <strong>and</strong> therefore outdoor recreation is common<br />

particularly in forests that are located near residential areas. In this article we define urban<br />

forests as forests with indigenous understorey vegetation. Thus, they do not include gardens,<br />

parks <strong>and</strong> street trees.<br />

Urban forestry has been defined on an European level as “planning, design, establishment <strong>and</strong><br />

management of trees <strong>and</strong> forest st<strong>and</strong>s with amenity values, situated in or near urban areas”<br />

(Forrest et al. 1999). In fact, urban forests provide multiple values for inhabitants. Therefore<br />

recreation, aesthetics <strong>and</strong> biodiversity have to be taken seriously into consideration when<br />

planning <strong>and</strong> managing these forests.<br />

Municipalities own most of the urban forests in Finl<strong>and</strong>. Participatory planning has been used<br />

for urban forest planning in municipalities since 1990’s (Löfström 1990, 1996, 2001). The<br />

purpose of the participatory planning is to gather information, wishes <strong>and</strong> preferences from local<br />

inhabitants <strong>and</strong> other users of urban forests to different planning processes that concern the<br />

future use of these forests. In its best, this kind of approach gives people an actual opportunity<br />

to influence the way the urban forests are managed in their surroundings. Participatory planning<br />

* Corresponding author. Tel.: +358 50 391 2418 - Fax: +358 50 391 2202<br />

Email address: irja.lofstrom@metla.fi<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Löfström et al. 2010. Biodiversity <strong>and</strong> recreational values in urban participatory forest planning in Finl<strong>and</strong><br />

611<br />

can also be used to solve local problems <strong>and</strong> even conflicts (Löfström 2006, Löfström etc 2007,<br />

2008a, 2000b, Mikkola etc 2008).<br />

The aim of the study was to gain an overall picture of the planning <strong>and</strong> management of<br />

municipally owned urban forests <strong>and</strong> to discover together with the urban practitioners the main<br />

development needs of planning processes. We studied especially if multiple values, e.g.<br />

recreation <strong>and</strong> biodiversity <strong>and</strong> forest users’ opinions were taken into account. Another aim was<br />

to study the applicability of multi-criteria decision analysis <strong>and</strong> advanced decision-support tools<br />

to urban forest planning.<br />

The aim of our case study in Puijo, located in the city of Kuopio, was to develop optimal<br />

participatory planning process <strong>and</strong> to investigate environmental values, views <strong>and</strong> opinions of<br />

citizens concerning forest management in Puijo.<br />

2. Methodology<br />

The research methods used were mail surveys <strong>and</strong> interviews of urban forest planners, working<br />

groups <strong>and</strong> seminars involving forest planners <strong>and</strong> other stakeholders as well as analysis <strong>and</strong><br />

follow-up of ongoing municipal forest management planning processes. This article consists of<br />

results of several studies which have been carried out during 2005-2009.<br />

In the recent Puijo planning process, the opinions of local people were collected with different<br />

methods. The data were collected in 2009 in the surrounding housing areas of Puijo forest<br />

through both mail survey <strong>and</strong> Internet survey during the participatory forest planning project.<br />

The mail survey was sent to 2000 inhabitants <strong>and</strong> about 25 % of them responded to the inquiry.<br />

3. Result<br />

The majority of the municipalities used lighter <strong>and</strong> a wider range of forest management methods<br />

for recreational forests than are generally in use in commercial forests. In recreational forests<br />

nature-oriented forest management methods were preferred. According to municipalities’ view<br />

old forest st<strong>and</strong>s, broad-leaved trees, <strong>and</strong> natural regeneration increase biodiversity, aesthetic<br />

<strong>and</strong> recreational values. However, municipalities have received also negative feedback from<br />

forest users when they have left decaying trees valuable for biodiversity into urban forests.<br />

Conflicts arising from enhancing both the biodiversity <strong>and</strong> the recreational use of forests may<br />

partly be due to the lack of knowledge. For example decaying wood has been collected for fire<br />

wood. Many forest users preferred forests where walking <strong>and</strong> running is easy, which indicates<br />

that decaying trees should be removed. In addition, safety of recreational forests is important for<br />

municipalities, <strong>and</strong> that has been one reason for removing st<strong>and</strong>ing dying trees. A solution<br />

might be to leave decaying trees into clusters that are not located near pathways.<br />

Municipalities are interested in improving forest biodiversity in the recreational forests they<br />

own. In their forests, many municipalities have actively inventoried the occurrence of<br />

ecologically valuable habitats <strong>and</strong> features important for, e.g. threatened <strong>and</strong> endangered species.<br />

However, municipalities expect more instructions <strong>and</strong> training on how to protect biodiversity in<br />

practice in their forests. They consider that financial incentives are also important for promoting<br />

biodiversity Compensation should be based on diminished returns <strong>and</strong> raised costs for forest<br />

management due to e.g. habitat restoration.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Löfström et al. 2010. Biodiversity <strong>and</strong> recreational values in urban participatory forest planning in Finl<strong>and</strong><br />

612<br />

Today, more than half of the municipalities use participatory planning for their forests. There<br />

are various approaches for participatory planning. Municipalities arrange meetings for the<br />

general public, invite inhabitants <strong>and</strong> stakeholders to take part in planning groups, or inquire the<br />

expectations <strong>and</strong> wishes of inhabitants by using questionnaires. The use of the Internet as a<br />

means of interaction has also increased. Some municipalities were not very satisfied with the<br />

achievements gained with public participation. One of the practical problems was the<br />

integration of various qualitative feedback truly into forest planning -the question is how to<br />

combine the different wishes of users <strong>and</strong> owners in the actual planning process. In addition<br />

clear objectives for public participation were often lacking.<br />

In Puijo's case the opinions of the users of Puijo <strong>and</strong> citizens living near the area were studied in<br />

the beginning of forest planning process <strong>and</strong> opinions were well taken into account e.g. in the<br />

creation of the plan alternatives. Citizens had different views concerning management methods<br />

that could be applicable to Puijo forest. Some citizens wanted more intensive management in<br />

future <strong>and</strong> accepted even clear cuttings in Puijo, although quite many of the citizens perceived<br />

most of all biodiversity <strong>and</strong> untouched Puijo forest in future. The study showed that many<br />

citizens had emotional relationship to Puijo <strong>and</strong> this makes the attitudes <strong>and</strong> opinions towards<br />

forest management stronger.<br />

The study results also indicate that most municipalities do not have clear objectives for planning<br />

<strong>and</strong> managing urban forests. Most of the municipalities had a management plan for their forests.<br />

In some municipalities, these plans covered only commercial forests. However, there was<br />

considerable variation in the planning st<strong>and</strong>ards <strong>and</strong> practices. In most cases, alternative forest<br />

plans <strong>and</strong> management schedules were missing. Roughly one fifth of the municipalities have<br />

drawn up a strategic urban forest plan for their forests, setting down principles for planning <strong>and</strong><br />

management for the next ten years.<br />

4. Discussion<br />

Nature-oriented forest management was preferred more in municipally owned recreational<br />

forests than in commercial forests. In fact, since the 1980’s, there has been significant shift in<br />

urban forest management practice towards the consideration of forest aesthetics, recreation <strong>and</strong><br />

biodiversity (Löfström 1990). The similar trend has occurred also in the management of urban<br />

forests in the other Nordic countries (Gundersen et al. 2005).<br />

Nature-oriented forest management was regarded as a way to increase both ecological <strong>and</strong><br />

aesthetic values. However, there were some contradictions between the enhancement of<br />

recreational use <strong>and</strong> the biodiversity of the forests. Participatory planning is a rather new tool<br />

for integrating aesthetic <strong>and</strong> ecological values for the management of urban forests. The use of<br />

this method for urban forest planning in municipalities has increased clearly: in the year 2006 it<br />

was three times more common than in 2000 (Löfström 2001, 2006). Thus, the participatory<br />

planning of municipally owned urban forests has become almost a rule in recent years in<br />

Finl<strong>and</strong>. Also a clear strategic viewpoint for owning forests is becoming more widespread. In<br />

the year 2000, only 6 % of municipalities had already completed a strategic urban forest plan for<br />

their forests, whereas the percentage in 2006 was about 20.<br />

Municipalities had actively inventoried biodiversity values of urban forests <strong>and</strong> involved local<br />

inhabitants <strong>and</strong> stakeholders into forest planning processes. However, the practical problems<br />

were e.g. the integration of various qualitative data into forest planning process <strong>and</strong> the lack of<br />

objectives for planning <strong>and</strong> management of forests <strong>and</strong> participation. These problems were<br />

partly due to insufficient resources <strong>and</strong> a lack of forest practitioners in municipalities.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Löfström et al. 2010. Biodiversity <strong>and</strong> recreational values in urban participatory forest planning in Finl<strong>and</strong><br />

613<br />

Participatory planning of Puijo forest was a challenging process when trying to integrate a range<br />

of different values into the management <strong>and</strong> to cope with the strong emotions, opinions <strong>and</strong><br />

expectations of the citizens. Multiple values, for example recreation, aesthetics <strong>and</strong> biodiversity<br />

have to be taken into consideration when managing Puijo forest in future.<br />

Multi-functional forest planning, multi-criteria decision analysis <strong>and</strong> advanced decision-support<br />

tools are rather new approaches in the field of urban forestry. However, this study showed that<br />

these modern methods could be useful tools for integrating multiple values – such as<br />

recreational, ecologic, cultural, aesthetic values- into planning process <strong>and</strong> thus to improve the<br />

quality of urban forest planning.<br />

References<br />

Forrest, M., R<strong>and</strong>rup, T.B., <strong>and</strong> Konijnendijk, C.-C. (Eds.) 1999. Urban forestry - research <strong>and</strong><br />

development in Europe. Report of COST Action E12 'Urban <strong>Forest</strong>s <strong>and</strong> Trees' on the state<br />

of the art of urban forestry research <strong>and</strong> development in Europe. European Union, COSTprogramme,<br />

Brussels, 363 p.<br />

Gundersen, V., Frivold, L.H., Löfström, I., Jorgensen, B.B., Falck, J. & Oyen, B.-H. 2005.<br />

Urban woodl<strong>and</strong> management - The case of 13 major Nordic cities. Urban <strong>Forest</strong>ry & Urban<br />

Greening, 3(3-4): 189-202.<br />

Löfström, I. 1990. Kaupunkien ja kuntien metsien hoito. Summary: The management of<br />

municipal forests. Ministry of Environment, Environmental Protection Department.<br />

Publication 87, 118 p.<br />

Löfström, I. 1996. The management of urban forests in Finl<strong>and</strong>. In: T.B. R<strong>and</strong>rup <strong>and</strong> K.<br />

Nilsson (Eds.). Urban <strong>Forest</strong>ry in the Nordic Countries. Danish <strong>Forest</strong> <strong>and</strong> l<strong>and</strong>scape<br />

Research Institute.<br />

Löfström, I. 2001. Taajamametsät suunnittelun kohteena. (In English: Planning of urban forests)<br />

In: Kangas, J. & Kokko, A. (Eds.). Metsän eri käyttömuotojen arvottaminen ja<br />

yhteensovittaminen. Metsän eri käyttömuotojen yhteensovittamisen tutkimusohjelman<br />

loppuraportti. Metsäntutkimuslaitoksen tiedonantoja 800: 260-261.<br />

Löfström, 2006. Municipalities keen on biodiversity in their forests. METSO Newsletter 4, 4 p.<br />

Löfström, Mikkola, N. & Tenhola, T. 2007. Kuntien virkistys- ja ulkoilumetsät. Julkaisussa:<br />

Syrjänen, K., Horne, P., Koskela, T. & Kumela, H. (toim). METSOn seuranta ja arviointi –<br />

Etelä-Suomen metsien monimuotoisuusohjelman seurannan ja arvioinnin loppuraportti.<br />

Maa- ja metsätalousministeriö, ympäristöministeriö, Metsäntutkimuslaitos ja Suomen<br />

ympäristökeskus, Vammala: 234–239.<br />

Löfström, Kurttila, M., Mikkola, N., Pykäläinen, J. & Tikkanen, J. 2008a. Multifunctional<br />

planning of municipally owned urban forests in Finl<strong>and</strong>. Julkaisussa: Sipilä, M., Tyrväinen,<br />

L. & Virtanen, E. (toim.). <strong>Forest</strong> Recreation & Tourism Serving Urbanized Societies. Joint<br />

Final Conference of <strong>Forest</strong> for Recreation <strong>and</strong> Tourism (COST E33) <strong>and</strong> 11th European<br />

Forum on Urban <strong>Forest</strong>ry (EFUF) 28.-31.5.2008, Hämeenlinna, Finl<strong>and</strong>. Abstracts. Finnish<br />

<strong>Forest</strong> Research Institute, Hansaprint Oy, Vantaa. p. 25.<br />

Löfström, Mikkola, N., Kurttila, M., Leskinen, P., Hujala, T., Jauhiainen, S. & Räsänen, J.<br />

2008b. Puijon metsäalueen hoidon ja käytön vuorovaikutteinen suunnittelu. Kuopion<br />

kaupunki, 27 p.<br />

Mikkola, N., Pykäläinen, J., Löfström, I., Kurttila, M. ja Tikkanen, J. 2008. Kuntametsien<br />

suunnittelun tiekartta -hankkeen loppuraportti. Metsäntutkimuslaitoksen työraportteja 68.<br />

http://www.metla.fi/julkaisut/workingpapers/2008/mwp068.htm<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.C.R. de Oliveira & S.M. Carvalho 2010. Urban tree inventory <strong>and</strong> socio-economic aspects of three villages of Ponta Grossa<br />

614<br />

Urban tree inventory <strong>and</strong> socio-economic aspects of three villages of<br />

Ponta Grossa, PR<br />

Ana Carolina Rodrigues de Oliveira * & Sílvia Méri Carvalho<br />

State University of Ponta Grossa, Ponta Grossa, PR, Brazil<br />

Abstract<br />

This study aimed at an analysis of urban tree road in villages Esmeralda, Jardim Carvalho <strong>and</strong><br />

Vilela, Bairro Jardim Carvalho in Ponta Grossa, identifying the percentage of native <strong>and</strong> exotic<br />

species <strong>and</strong> also relate to afforestation with the socio-economic aspects of the local population.<br />

It also seeks to provide subsidies to the municipal government for the development of an<br />

afforestation plan. Measurements were taken of the sidewalks, as well as analysis of possible<br />

conflicts with the space that is. About individual trees, 27% are native species <strong>and</strong> 73% exotic,<br />

<strong>and</strong> the species Lagerstroemia indica L. (extremosa) the species that stood out with 18%. This<br />

result demonstrates the trend in urban areas the prevalence of exotic species <strong>and</strong> the weak<br />

participation of the green element in the urban l<strong>and</strong>scape. Were identified conflicts with urban<br />

facilities highlighting <strong>and</strong> the need for a correct <strong>and</strong> efficient management.<br />

Key-words: Urban Afforestation, Exotic, Native<br />

1. Introduction<br />

The addition of trees to the ecosystem elements anthropogenically modified, typical of<br />

the environment of cities (Kulchetsck, 2006), brings countless benefits. Mascaró <strong>and</strong> Mascaró<br />

(2002) propose, therefore, trees in cities, highlighting the important role that vegetation is as<br />

soothing urban pollution, in addition to environmental, energy <strong>and</strong> l<strong>and</strong>scape. Soon, vegetation,<br />

justified also by chemical factors, physical, ecological <strong>and</strong> psychological, should be considered<br />

in the order of the priorities of urban planning. Search is thus a return to lost balance <strong>and</strong> a<br />

significant improvement in quality of life.<br />

Factors that should be considered in urban trees are many <strong>and</strong> can be highlighted: the<br />

urban environment, characterized in terms of climate, soils, topography, the physical space<br />

available in relation to the width of streets <strong>and</strong> sidewalks, l<strong>and</strong> clearance, the height the<br />

buildings <strong>and</strong> the presence of electrical cords air, water pipe, sewer, storm sewers, telephone<br />

network, the characteristics of the species to be used, in regard to climatic adaptability,<br />

resistance to pests <strong>and</strong> diseases, tolerance to pollution, lack of principles toxic or allergenic, <strong>and</strong><br />

phenological characteristics (shape, size, root, flowering, fruiting, etc.) <strong>and</strong> morphological<br />

characteristics. Examined the issue of ecological adaptation (acclimatization, naturalization,<br />

housing) should be alert to the readiness of the city for the trees, not introducing the species to<br />

be r<strong>and</strong>om. (Urban Tree, 2004; Sampaio, 2006; Serafim, 2007).<br />

This work is part of a larger project that aims to inventory all city trees, providing support<br />

to the government to implement public policies related to this topic.<br />

* Corresponding author. Tel.: (55) 42 3225-9206<br />

E-mail: krowmanson@hotmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.C.R. de Oliveira & S.M. Carvalho 2010. Urban tree inventory <strong>and</strong> socio-economic aspects of three villages of Ponta Grossa<br />

615<br />

This research aims to analyze the current context of the afforestation of public roads from<br />

villages Esmeralda, Jardim Carvalho <strong>and</strong> Vilela in Ponta Grossa, identifying native <strong>and</strong> exotic<br />

species present, <strong>and</strong> also examine the afforestation as a reflection of the socio- economic aspects<br />

of population.<br />

2. Characterization of study area<br />

Ponta Grossa, located under the average elevation of 975 meters, is included in the<br />

Second Paraná Plateau in the east-central region of Parana State, being an important road <strong>and</strong><br />

rail junction of the state. (IBGE, 2008).<br />

The Jardim Carvalho is located in the northeast of the city <strong>and</strong> houses within its limits<br />

thirteen villages. Were chosen as the spatial area of the district three villages, namely:<br />

Esmeralda, Jardim Carvalho <strong>and</strong> Vilela. We opted for these three towns they are<br />

demographically more concentrated <strong>and</strong> have vast socio-economic disparities, <strong>and</strong> were the<br />

closest to the downtown area.<br />

The three villages comprising 11 census tracts, which total, according to IBGE (2000),<br />

6971 inhabitants, with 2007 inhabitants in the village Jardim Carvalho, 602 in the village<br />

Esmeralda <strong>and</strong> 4362 in Vilela.<br />

As one of the objectives of the research was to investigate the relationship of urban<br />

environmental quality (through the indicator of urban forestry) with socioeconomic indices, data<br />

from the 2000 census were prevalent for the analysis. We selected some variables that<br />

characterize the people responsible for permanent private homes - such as education <strong>and</strong> income<br />

- <strong>and</strong> to those which have stressed the social inequality between the villages Esmeralda / Jardim<br />

Carvalho <strong>and</strong> Vilela.<br />

Looking at Figure 1 it is noted that the rate of people with lower educational level is<br />

higher in Vilela, <strong>and</strong> 2% of the illiterate population, while in villages Esmeralda <strong>and</strong> Jardim<br />

Carvalho that number drops to 0.3 <strong>and</strong> 0.2, respectively . The variable 'Course highest attended -<br />

College' was the one that showed the difference between the two groups this census: Jardim<br />

Carvalho <strong>and</strong> Esmeralda has more than twice as many people with this level of education<br />

compared with Vilela. It should be noted that this village has 1,700 people more than the other<br />

two villages, thus demonstrating how low is that index.<br />

Thus, from the variables ‘Course highest attended – College’ <strong>and</strong> '15 years of study' is<br />

possible to conclude that the villages Esmeralda <strong>and</strong> Jardim Carvalho have better rates of<br />

schooling than Vilela.<br />

People<br />

(%)<br />

12<br />

9<br />

6<br />

3<br />

0<br />

2<br />

0,2 0,3 0,4 0,3<br />

Illiterate<br />

2<br />

Course highest<br />

attended - no<br />

course<br />

12<br />

13<br />

3<br />

Course highest<br />

attended -<br />

college<br />

5 5<br />

1<br />

15 years of study<br />

Jardim Carvalho<br />

Esmeralda<br />

Vilela<br />

Figure 1 - Schooling of heads of permanent private households in villages Esmeralda, Jardim Carvalho<br />

<strong>and</strong> Vilela. Source: IBGE (2000).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.C.R. de Oliveira & S.M. Carvalho 2010. Urban tree inventory <strong>and</strong> socio-economic aspects of three villages of Ponta Grossa<br />

616<br />

In relation to local income, which can be said, reflects the level of education of the<br />

population is concentrated in villages Esmeralda <strong>and</strong> Jardim Carvalho. One sees this phenomenon<br />

in Figure 2, which illustrates, also in proportion to the population, these data. The greater the<br />

number of monthly income of heads of households, increased economic inequality, because these<br />

two villages are increasing numbers with increasing income, unlike Vilela, which the numbers<br />

decrease with the increase of it. This village puts forward in the two graphs were examined, the<br />

worse indices of education <strong>and</strong> monthly income, thus setting a framework for social <strong>and</strong><br />

economic disparities between villages so close.<br />

(%)<br />

People<br />

9<br />

6<br />

3<br />

0<br />

2<br />

0,6 0,6<br />

Without income<br />

6<br />

5<br />

5<br />

4<br />

2 2 2<br />

1 1<br />

1 to 2 minimum<br />

wages<br />

10 to 15<br />

minimum wages<br />

more than 20<br />

minimum wages<br />

Jardim Carvalho<br />

Esmeralda<br />

Vilela<br />

Figure 2– Nominal monthly income of heads of permanent private households in villages Esmeralda,<br />

Jardim Carvalho <strong>and</strong> Vilela. Source: IBGE (2000).<br />

From the data presented was based on the hypothesis that these socioeconomic indices<br />

may reflect quantitative <strong>and</strong> qualitative urban forestry in this area, a phenomenon that will be<br />

addressed before the data obtained from field work.<br />

3. Methodology<br />

We first performed a literature on the subject trees, especially of roads, thereby having<br />

contact with the methodologies already applied. Were also selected data from Census 2000 to<br />

assess the environmental quality while reflecting socio-economic development.<br />

For delimitation of the study area, as well as research planning, cartograms were produced<br />

to map the paths to be scheduled through the software Arc View 3.2 GIS Laboratory,<br />

Department of Geosciences, State University of Ponta Grossa. Were used as the digital<br />

cartographic base of the municipality of Ponta Grossa.<br />

The following worksheets were produced for species identification, linking individual<br />

trees in each side of the road (right / left) <strong>and</strong> the distances in which they were in the building<br />

<strong>and</strong> the curb, apart from possible conflicts with the structures the city (like breaking sidewalks<br />

<strong>and</strong> conflict with the electric grid).<br />

Such information has been verified through research on site, which it based individuals<br />

with PBH (perimeter at breast height), less than 20 cm, <strong>and</strong> they were clearly located on<br />

sidewalks <strong>and</strong> public walkways. When the impossibility of identification in the field, samples<br />

were collected for identification in Herbarium of State University of Ponta Grossa with the aid<br />

of works Lorenzi <strong>and</strong> Souza (1999), Lorenzi (2002) <strong>and</strong> Lorenzi et al (2003).<br />

In the research field were also made photographic records of the species most frequent<br />

conflicts <strong>and</strong> ways identified with potential for afforestation. Finally, with the data already<br />

collected, comparisons were made with other sites of Ponta Grossa already inventoried, to have<br />

a diagnosis of urban forestry roads in the area.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.C.R. de Oliveira & S.M. Carvalho 2010. Urban tree inventory <strong>and</strong> socio-economic aspects of three villages of Ponta Grossa<br />

617<br />

4. Result <strong>and</strong> Discussion<br />

We covered 59 routes, which had 479 individual trees to the total. Of these, 29 were<br />

identified at family level, 19 at genus level <strong>and</strong> 411 species, of which 27% are natives <strong>and</strong> 73%<br />

exotic. Due to the absence of flowers <strong>and</strong> / or fruit (elements necessary for correct<br />

identification) <strong>and</strong> the radical pruning, 20 tree specimens were not identified.<br />

The species most often was the Lagerstroemia indica L., commonly known as the<br />

Extremosa, the family Lythraceae, constituting 18% of total species (74 individuals).<br />

The second species was the most present Ligustrum lucidum, Oleaceae family, with<br />

15.32% followed by the species of Ficus benjamina Moraceae with 10.94%. It is noteworthy<br />

that the species Ligustrum Lucidum had many conflicts with sidewalks, like Ficus benjamina,<br />

which, because of its shallow root is not recommended for narrow forest roads. It is therefore<br />

important to know the characteristics of the species for its correct use. It is also recommended to<br />

research the species native to the area thereby avoiding the excessive use of exotic species.<br />

For Santamur Junior (2002, apud Quadros, 2005) does not exceed more than 10% of the<br />

same species, 20% of the same genus <strong>and</strong> 30% from the same botanical family. According to this<br />

author, the species Lagerstroemia indica <strong>and</strong> Ligustrum lucidum exceeded the limit of tree<br />

species recommended for health (18% <strong>and</strong> 15.32%, respectively). In relation to any botanical<br />

families exceeded this parameter, the most frequent: Bignoniaceae with 16%, 15% Lythraceae,<br />

Fabaceae with 14% <strong>and</strong> 13% to Oleaceae.<br />

As surveyed, there are only 39 species in the arborization of the villages Esmeralda, Jardim<br />

Carvalho <strong>and</strong> Vilela, predominantly exotic on the native. The three most common species<br />

together represent 44.26% of the total cataloged.<br />

The average distance from the curb found for the sampled population was 0.52 m. Value<br />

lower than that found in other locations as those raised by Milano (1984, 1988 apud Loboda et al<br />

2005) <strong>and</strong> Ng (1995 apud Loboda et al 2005), from 1.56 to Curitiba / PR, <strong>and</strong> 1.20 m Maringá /<br />

PR , <strong>and</strong> 2.1 m for Cascavel / PR, respectively. The average distance from buildings (1.20 m) was<br />

also a low rate when compared with the values found by Nunes (1995 apud Loboda et al 2005) in<br />

Apucarana <strong>and</strong> Cascavel, respectively, 2.41 I 3.3 m, <strong>and</strong> Maringá, 1.47 m (ibid.). Data concerning<br />

the average distance of trees to the curb <strong>and</strong> the buildings show an average width of 1.72 m for<br />

the tours of the area sampled. Therefore, tours of small, should include trees, shrubs <strong>and</strong> small<br />

trees, since the limited space hinders the development of afforestation.<br />

The lack of planning of afforestation culminates in conflicts with urban facilities such as<br />

the presence of spinning air, which is one of the most important factors when planning the<br />

arborization. It was observed conflicts with the electric grid in 68 cases (28.57%), often having to<br />

be done pruning, changing the natural shape of the tree also produces an anti-aesthetic.<br />

The lack of free area <strong>and</strong> the choice for species with shallow root system eventually<br />

undermine, among other urban facilities, the sidewalks, which end up <strong>and</strong> breaking. Therefore<br />

one should choose a tree with deep roots, <strong>and</strong> leave at least 1m ² of space pavement that allow the<br />

infiltration of water <strong>and</strong> nutrients (Santos <strong>and</strong> Teixeira, 2001), avoiding situations like breaking<br />

sidewalks, which represent 148 cases (62.18%) found in the study area.<br />

Another practice is deeply rooted in Brazil's painting trunks, <strong>and</strong> found 22 cases in the<br />

sampled area. For Santos <strong>and</strong> Teixeira (2001) this practice provides dubious aesthetic effect <strong>and</strong><br />

can cause damage to health, as the bark of trees has its own defenses.<br />

Among the 479 tree specimens found, 389 are located in the Jardim Carvalho <strong>and</strong><br />

Esmeralda Village, in other words, 81%, leaving only 90 for Vilela, or 19%. Therefore, making<br />

a correlation, one realizes that the urban environmental quality also reflects income inequality,<br />

as stated Berto (2008). With the socio-economic data from the IBGE this paper, the social<br />

disparity between the villages Esmeralda / Jardim Carvalho <strong>and</strong> Vilela was evident, with the<br />

indices of trees obtained confirmed the importance that the income has in the establishment of<br />

environmental quality. Contributed to this analysis the fact the town Vilela have the lowest<br />

levels of education, income <strong>and</strong> only 19% of trees cataloged in the search.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.C.R. de Oliveira & S.M. Carvalho 2010. Urban tree inventory <strong>and</strong> socio-economic aspects of three villages of Ponta Grossa<br />

618<br />

We agree with Seraphim (2007, p. 4) when it concludes that "aspects of environmental<br />

quality may be clear in some localities in urban areas depending on the distribution of<br />

vegetation <strong>and</strong> indicate the quality of life of residents."<br />

In the field research were also identified five pathways (Adjaniro Cardon, Graciliano<br />

Ramos, Rocha Pombo, Monte Alverne <strong>and</strong> Henrique Thielen) who all have potential to be<br />

wooded. In these, the hunt is wider than 3m, the streets are wide, the houses are setback <strong>and</strong><br />

spinning air is absent.<br />

Conclusion<br />

It was evident that the urban environmental quality also reflects the income inequality<br />

since the villages with better socio-economic - <strong>and</strong> therefore more assisted by basic<br />

infrastructure - they also presented a greater number of trees <strong>and</strong> a greater awareness about<br />

environmental issues. So we can conclude that education is strongly related to a condition where<br />

the theme Environmental urban areas.<br />

Through the survey <strong>and</strong> analysis of arborization, we could perceive that there is an<br />

unequal distribution of individual trees, focusing on villages Esmeralda <strong>and</strong> Jardim Carvalho<br />

with 81% of the tree, against only 19% for the village Vilela.<br />

Encouragement of the arborization should be taken, considering that seven streets were<br />

not found any tree, <strong>and</strong> there are roads with trees <strong>and</strong> very few individuals with the potential to<br />

make plantations without risk of conflict with the sidewalk or spinning air.<br />

References<br />

Berto, V. Z. 2008. Análise da Qualidade Ambiental Urbana na Cidade de Ponta Grossa (PR):<br />

Avaliação de algumas propostas metodológicas. Ponta Grossa, Paraná. 150p.<br />

Gallina, M. H. Verona, J. A; Troppmair, H. 2003 Geografia e questões ambientais. Mercator, n.<br />

4, p. 87-97,<br />

IBGE – 2009. Census Data 2000. Avaliable in: www.ibge.gov.br, accessed: August 22.<br />

Loboda, C. R.; De Angelis, B. L. D. 2005 Áreas Verdes Públicas Urbanas: Conceitos, Usos e<br />

Funções. Ambiência Magazine – Pag. 125 to 138.<br />

Lorenzi, H.; Souza, H. M (de).1999. Plantas Ornamentais no Brasil: arbustivas, herbáceas e<br />

trepadeiras. Nova Odessa: Instituto Plantarum, 1088 p.<br />

Lorenzi, H. 2002. Árvores brasileiras: manual de identificação e cultivo de plantas arbóreas do<br />

Brasil. Nova Odessa: Instituto Plantarum, 384p.<br />

Lorenzi, H.; Souza, H.M. De.; Torres, M.A.V.; Bacher, L.B. 2003. Árvores exóticas no Brasil:<br />

madeiras, ornamentais e aromáticas. Nova Odessa: Instituto Plantarum, 368p.<br />

Mascaró, L.; Mascaró, J.2005. Vegetação urbana. Porto Alegre, 204p.<br />

Mir<strong>and</strong>a, T. O. 2008. Arborização Urbana Viária no Bairro da Ronda, Ponta Grossa - PR:<br />

Composição e Avaliação. Ponta Grossa -PR, 95 p.<br />

Quadros, G. P. 2005. Arborização Urbana na Área Central de Ponta Grossa: Implantação,<br />

Preservação e Monitoramento - 2005. Ponta Grossa. 102 p.<br />

Sampaio, A. C. F. 2006. Análise da Arborização de Vias Públicas das Principais Zonas do<br />

Plano Piloto de Maringá-PR. Maringá-PR. 117p.<br />

Santos, N. R. Z.; Teixeira, I. F. 2001. Arborização de Vias Públicas: Ambiente x Vegetação. RS:<br />

Clube da árvore. 135 p.<br />

Serafim, A.R.M.D.B. 2007. O verde na cidade: análise da cobertura vegetal nos Bairros do<br />

centro exp<strong>and</strong>ido da cidade do Recife/PE. Universia.11 p.<br />

Silva, R.K.D. 2006. Arborização Urbana Viária no Bairro de Olarias, Ponta Grossa/PR. Ponta<br />

Grossa, Paraná. 75 p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A.C.R. de Oliveira & S.M. Carvalho 2010. Urban tree inventory <strong>and</strong> socio-economic aspects of three villages of Ponta Grossa<br />

619<br />

Vilela, J. C. 2007. Levantamento Quantitativo e Qualitativo de Individuos Arbóreos Presentes<br />

nas vias do Bairro Estrela em Ponta Grossa/PR. Ponta Grossa, Paraná. 96p.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Silva et al. 2010. Lisbon’s public gardens, host place for world’s trees<br />

620<br />

Lisbon’s public gardens, host place for world’s trees<br />

Isabel Silva 1 , Elsa Isidro 1 , Ana Luísa Soares 2* & Francisco Moreira 2<br />

1<br />

Secção Autónoma de Arquitectura Paisagista, Instituto <strong>Superior</strong> de Agronomia,<br />

Universidade Técnica de Lisboa, Portugal<br />

2<br />

Centro de Ecologia Aplicada Prof. Baeta Neves, Instituto <strong>Superior</strong> de Agronomia,<br />

Universidade Técnica de Lisboa, Portugal<br />

Abstract<br />

This study aims to contribute to the characterization <strong>and</strong> evaluation of Lisbon’s Gardens. Within<br />

the framework of the 2009 project “Methods of Characterization <strong>and</strong> Classification of Lisbon’s<br />

Public Gardens with heritage interest”, 31 of Lisbon’s Public Gardens were studied <strong>and</strong> a new<br />

methodology was developed to measure their l<strong>and</strong>scape, historical, social <strong>and</strong> cultural value<br />

Garden’s L<strong>and</strong>scape value was evaluated according to several parameters, one of which was the<br />

botanical quality indicator. It determines trees’ richness <strong>and</strong> uniqueness, assessed by botanical<br />

diversity <strong>and</strong> singularity evaluation methods (e.g., surveys, Shannon index, Equitability), <strong>and</strong><br />

trees’ heritage interest, based on their rarity, age, size <strong>and</strong> health.<br />

3751 trees in 31gardens were studied. The following results were obtained: 46 families, 83<br />

genus <strong>and</strong> 139 species. 48% are exotic, 35% naturalized, 13% indigenous <strong>and</strong> 4% exotic<br />

invaders.<br />

Lisbon’s Mediterranean climate allows the coexistence of different tree species, from Northern<br />

Europe to subtropical climates. In addition to its aesthetical value, this botanical diversity plays<br />

a central role in increasing biodiversity <strong>and</strong> promoting urban ecological sustainability.<br />

Keywords: Public Gardens, Heritage Interest, garden’s trees, Botanical Diversity, Lisbon<br />

1. Introduction<br />

Lisbon’s climate allows for the coexistence of various indigenous <strong>and</strong> exotic tree species, from<br />

Northern Europe to subtropical climes. In addition to its great aesthetical value, this botanical<br />

diversity provides a habitat for the fauna <strong>and</strong> plays a key role in increasing biodiversity <strong>and</strong> in<br />

the urban ecology thus contributing to a sustainable city.<br />

This botanical richness is mostly due to the Portuguese Discoveries <strong>and</strong> contact with other<br />

cultures which meant that plant species from around the world were brought to Portugal, in<br />

particularly to Lisbon. These “new” plants, from all over the world, were a challenge for<br />

botanists, gardeners <strong>and</strong> horticulturists, for whom the public <strong>and</strong> private botanical gardens <strong>and</strong><br />

gardens were the “stage” for their experiments.<br />

Many of these species were well suited to our climate, <strong>and</strong> today in Lisbon’s streets, parks <strong>and</strong><br />

gardens we can find specimens such as: Tipuana tipu, Phytollaca dioica, Chorisia speciosa <strong>and</strong><br />

Jacar<strong>and</strong>a mimosifolia from South America; palm trees from the Canaries; Casuarina<br />

cunninghamiana, Grevillea robusta <strong>and</strong> Lagunaria patersonii from Australia, Taxodium<br />

distichum from the USA, Metrosideros excelsa from New Zeal<strong>and</strong>, alongside Portuguese flora.<br />

Some of these trees, which st<strong>and</strong> out because of their size, structure, age, rarity or for historic<br />

<strong>and</strong> cultural reasons, have been classified by the National <strong>Forest</strong>ry Authority, <strong>and</strong> add to<br />

* Corresponding author. Tel.: 00 351 93 232 60 04<br />

Email address: alsoares@isa.utl.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Silva et al. 2010. Lisbon’s public gardens, host place for world’s trees<br />

621<br />

Lisbon’s ecological, l<strong>and</strong>scape, cultural <strong>and</strong> historic heritage. At the present time Lisbon’s tree<br />

collection is made up of 18 clusters, 54 single trees belonging to Lisbon Municipal Council <strong>and</strong><br />

6 privately owned single trees classified as trees <strong>and</strong>/or clusters of Public Interest.<br />

Under the Project “Methods of Characterization <strong>and</strong> Classification of Lisbon’s Public Gardens<br />

of Heritage Interest” (2009) a study was carried out to apply a new methodology to the<br />

measurement of the gardens’ L<strong>and</strong>scape, Historical <strong>and</strong> Socio-cultural value.<br />

The aim of this study is to contribute to the effective characterization <strong>and</strong> evaluation of Lisbon’s<br />

gardens. The overall value attributed to each garden was directly related to its heritage<br />

importance, use as public space <strong>and</strong> other factors.<br />

2. Methodology<br />

Under the Project “Methods of Characterization <strong>and</strong> Classification of Lisbon’s Public Gardens<br />

of Heritage Interest” † , a method was devised that gauges the Historic, L<strong>and</strong>scape <strong>and</strong><br />

Sociocultural Value of the 31 gardens studied (see Table 1). Parametric analysis of each value,<br />

using various quality indicators <strong>and</strong> a scoring system, made it possible to quantify each value<br />

<strong>and</strong> rank Public Gardens of Heritage Interest. The methodology developed <strong>and</strong> tested by the<br />

authors is based, among other principles, on surveys carried out by the National Trust of<br />

Engl<strong>and</strong> <strong>and</strong> English Heritage.<br />

The Historical value of each garden was evaluated by L<strong>and</strong>scape Architects, according to<br />

several indicators: origin, evolution, historic style, architectural <strong>and</strong> artistic integrity.<br />

Public surveys were conducted to determine the Socio-cultural value of the gardens role in<br />

creating healthy communities through recreation <strong>and</strong> economic sustainability (tourism), leading<br />

ultimately to better management.<br />

Table 1 - Methods of Characterization <strong>and</strong> Classification of Lisbon’s Public Gardens of Heritage Interest<br />

† In 2009, under an agreement between the Institute of Agronomy (Technical University of Lisbon) <strong>and</strong> Lisbon Municipal Council,<br />

final year L<strong>and</strong>scape Architecture students, Elsa Isidro <strong>and</strong> Isabel Silva, under the supervision of Professors Ana Luísa Soares <strong>and</strong><br />

Cristina Castel-Branco <strong>and</strong> the L<strong>and</strong>scape Architect Mafalda Farmhouse, developed a Method of Characterization <strong>and</strong> Classification<br />

of Lisbon’s Public Gardens of Heritage Interest, as part of the end of course work.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Silva et al. 2010. Lisbon’s public gardens, host place for world’s trees<br />

622<br />

In this paper the authors have created a broad scope concept that covers all the garden<br />

components which have been consciously designed to create a piece of L<strong>and</strong>scape art. This<br />

“L<strong>and</strong>scape Value” concept is arrived at through the evaluation of various elements such as<br />

artistic quality, aesthetic beauty, vegetation <strong>and</strong> l<strong>and</strong>scape, <strong>and</strong> built elements.<br />

The L<strong>and</strong>scape value<br />

L<strong>and</strong>scape value is a broad scope concept that includes all the components under the influence<br />

of the garden’s author in a process that leads to a piece of L<strong>and</strong>scape art, through the evaluation<br />

of the aesthetic beauty, artistic quality <strong>and</strong> value of the built elements or vegetation. This value<br />

includes also the garden’s external characteristics, such as the socioeconomic context of the<br />

garden’s location which acts as a parameter restricting the garden’s public to a certain kind of<br />

users, the aesthetic <strong>and</strong> artistic quality of the urban fabric, where we consider the state of<br />

conservation, <strong>and</strong>, finally, the visual relation between the garden <strong>and</strong> the surroundings, through<br />

a study of the series of views.<br />

As regards L<strong>and</strong>scape Value, which is a broad item, features were assessed such as the aesthetic<br />

<strong>and</strong> artistic quality of the built <strong>and</strong> botanical composition elements. As for botanical quality,<br />

following a survey ‡ of the tree species in the garden, their diversity <strong>and</strong> singularity were<br />

evaluated by means of specific scientific formulae. Since plants are a fragile element, easily lost,<br />

altered or replaced over time, it was also important to value those trees which because of their<br />

size, rarity or quality are deemed notable <strong>and</strong> classified as being of Public Interest § . In addition<br />

to these criteria the garden’s scenic quality was assessed by studying its plant composition in<br />

terms of trees, shrubs <strong>and</strong> flowers.<br />

Botanical quality pertains to the vegetation’s richness, uniqueness <strong>and</strong> heritage interest. The<br />

latter is based on each individual specimen’s size, rarity, age <strong>and</strong> health. The richness <strong>and</strong><br />

uniqueness of each specimen was obtained through accepted methods, as explained below.<br />

The Botanical Diversity parameter for each Garden was quantified using a average of four<br />

formulae: the Shannon index (1), the evenness index (2), the proportion of different species in<br />

the garden relative to the total number of specimens, <strong>and</strong> the ratio of the number of species in<br />

the garden compared to the total number of species in the set of gardens studied.<br />

Botanical Diversity Method:<br />

• Index 1: (Total of the tree species/ total of the garden’s tree specimens) (%)<br />

• a) Class 1 Botanical Diversity Index: (1 if [0-33%]; 2 if [34-66%]; 3 if [67-<br />

100%])<br />

• b) Class 1 Botanical Diversity Index: (1 if [0-15%]; 2 if [16-31%]; 3 if [32-<br />

50%])<br />

• Index 2: (Total of the tree species/Total of tree specimens in the set of gardens) (%)<br />

• a) Class 2 Botanical Diversity Index: (1 if [0-0,6%]; 2 if [0,7-1,2%]; 3 if [1,3-<br />

2%])<br />

• Index 3: Shannon index (1)<br />

• Class 3 Shannon Index: (1 if [0-1,33]; 2 if [1,34-2,66]; 3 if [2,67-4])<br />

• Index 4: Evenness (2)<br />

‡ The tree survey was conducted during the academic year 2008/2009 with the assistance of third year L<strong>and</strong>scape Architecture<br />

students at the Institute of Agronomy, Technical University of Lisbon.<br />

§ The classification of trees or groups of trees of Public Interest is governed by Decree-Law nº 20 985 of 1932-03-07 <strong>and</strong> Decree-<br />

Law nº 28 468 of 1938-02-15. The Ministry of Agriculture, Rural Development <strong>and</strong> Fisheries, through the AFN (National <strong>Forest</strong>ry<br />

Authority), is responsible for the classification pursuant to the terms of Regulatory Decree 10/2007 of 2007-02-27, published in the<br />

Official Journal (Diário da República) nº 41.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Silva et al. 2010. Lisbon’s public gardens, host place for world’s trees<br />

623<br />

• Class 4 Evenness Index (1 if [0-0,33]; 2 if [0,34-0,66]; 3 if [0,67-1])<br />

• Average of Botanical Diversity Indices<br />

ni : number of individuals in each species;<br />

abundance of each species.<br />

S : number of species (richness)<br />

N : total number of all individuals:<br />

p i : relative abundance of each species,<br />

calculated as the proportion of individuals of<br />

one species to the total number of individuals in<br />

the community:<br />

The Botanical Singularity score was based on the occurrence of one, two or three individuals of<br />

a particular species in the set of gardens studied.<br />

Singularity Evaluation Method:<br />

• 3 – contains the only specimen of a species;<br />

• 2 – contains one/two of the only two specimens of a species;<br />

• 1 - contains one/two/three of the only three specimens of a species;<br />

• 0 - if there are more than three specimens of a species<br />

The characterization of 31 public gardens according to the defined indicators enabled the<br />

acknowledgement of their social, historical <strong>and</strong> aesthetic features, <strong>and</strong> their ranking.<br />

3. Result <strong>and</strong> discussion<br />

A total of 3751 trees were recorded in the 31 gardens studied, <strong>and</strong> the following results<br />

were obtained: 46 families, 83 genus <strong>and</strong> 139 different species, of which 58% are<br />

evergreen <strong>and</strong> 42% deciduous. As regards their origin, 48% are exotic, 35% naturalized,<br />

13% indigenous <strong>and</strong> 4% exotic invaders. The dominant species are Celtis australis,<br />

Olea europaea var. sylvestris, Pinus pinea <strong>and</strong> Phoenix canariensis (see figure 1). The<br />

following species st<strong>and</strong> out because of their singularity: Erythrina crista-gallis,<br />

Firminiana simplex, Koelreuteria paniculata, Melaleuca stypheliodes <strong>and</strong> Pawlonia<br />

tomentosa.<br />

Figure 1 - Trees in Lisbon's Public Gardens<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Silva et al. 2010. Lisbon’s public gardens, host place for world’s trees<br />

624<br />

As for the gardens, the following are of special note due to their botanical diversity (see<br />

figure 2): Estrela (Guerra Junqueiro), Príncipe Real, Campo de Santana (Braamcamp<br />

Freire), Amoreiras (Marcelino Mesquita), <strong>and</strong> Vasco da Gama. The least diversity was<br />

found in São Pedro de Alcântara (António Nobre) <strong>and</strong> Campo Pequeno (Marquês de<br />

Marialva).<br />

Figure 2 – Botanical diversity in Lisbon’s Public Gardens<br />

Singularity analysis (see figure 3) showed that most gardens had unique specimens for<br />

different species, thus each garden on its own contributes to the overall diversity of trees<br />

present.<br />

Figure 3 – Botanical Singularity in Lisbon's Public Gardens<br />

4. Discussion<br />

Lisbon’s Mediterranean climate allows for the coexistence of different tree species, in addition<br />

to the gardens aesthetic value, this diversity is key to creating a healthy urban environment, with<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


I. Silva et al. 2010. Lisbon’s public gardens, host place for world’s trees<br />

625<br />

increasing biodiversity <strong>and</strong> sustainability. Besides their historic <strong>and</strong> cultural value, the public<br />

gardens studied are showplaces for plants from all over the world <strong>and</strong> bring an aesthetic value<br />

<strong>and</strong> high bio-climatic comfort to the city, as well as contribute to biodiversity. They have an<br />

unquestionable tourist <strong>and</strong> recreational value.<br />

References<br />

Soares, A.L. <strong>and</strong> Castel-Branco, C., 2007. As Árvores da Cidade de Lisboa. In: SILVA, J.S.<br />

(ed.). Floresta e Sociedade, uma história em comum. Público/FLAD/LPN. Lisboa: 289-<br />

334.<br />

Isidro, E., 2009. Metodologia de Caracterização e Classificação de Jardins Públicos de<br />

Interesse Patrimonial – Aplicação à cidade de Lisboa. Trabalho de Fim de Curso.<br />

Instituto <strong>Superior</strong> de Agronomia, Universidade Técnica de Lisboa, Lisboa (uupublished),<br />

121p.<br />

Silva, A. I., Isidro, E., Soares, A. L. <strong>and</strong> Castel-Branco, C., 2009. O Interesse Patrimonial dos<br />

Jardins Públicos de Lisboa. Actas de Comunicações do 6º Congresso Ibero-Americano de<br />

Parques e Jardins Públicos, Póvoa de Lanhoso: 21-30.<br />

Isidro, E.; Silva, I. <strong>and</strong> Soares, A.L., 2009. Jardim Braancamp Freire: diversidade botânica em<br />

colina histórica. Jardins 84: 10 - 15.<br />

Isidro, E.; Silva, I. <strong>and</strong> Soares, A.L., 2009. Jardim do Torel: um anfiteatro verde com vista sobre<br />

a cidade. Jardins 85: 12 - 15.<br />

Isidro, E.; Silva, I. <strong>and</strong> Soares, A.L., 2009. Jardim das Amoreiras: o jardim que promoveu a<br />

indústria da seda. Jardins 86: 18 - 21.<br />

Silva, I., Isidro, E. <strong>and</strong> Soares, A.L., 2009. Jardim Amália Rodrigues: um anfiteatro com<br />

múltiplos ambientes. Jardins 87: 16 – 19.<br />

Silva, I., Isidro, E. <strong>and</strong> Soares, A.L., 2010. Alameda D. Afonso Henriques: o eixo verde<br />

monumental do Estado Novo. Jardins 88: 12 – 15.<br />

Silva, I., Isidro, E. <strong>and</strong> Soares, A.L., 2010. Praça do Império: o icone verde da exposição do<br />

mundo português. Jardins 89: 10 – 14.<br />

Silva, I., Isidro, E. <strong>and</strong> Soares, A.L., 2010. São Pedro de Alcântara: um ícone da Lisboa<br />

Romântica do século XIX. Jardins 90: 14 – 17.<br />

Silva, I., Isidro, E. <strong>and</strong> Soares, A.L., 2010. Jardim 9 de Abril: homenagem à presença de<br />

Portugal na 1ª guerra mundial. Jardins 91: 16 – 19.<br />

Isidro, E.; Silva, I.. <strong>and</strong> Soares, A.L., 2009. Jardim do Príncipe Real: estilo romântico em<br />

ambiente histórico. Jardins 92: 16 - 19.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Section 9<br />

Symposia


Quantifying the effects of forest fragmentation: implications<br />

for l<strong>and</strong>scape planners <strong>and</strong> resource managers


Z.L. Urech & J.P. Sorg 2010. Taking into account local people’s livelihood systems for a better management of fragments<br />

628<br />

Taking into account local people’s livelihood systems for a better<br />

management of forest fragments<br />

Zora Lea Urech * & Jean-Pierre Sorg<br />

Groupe de foresterie pour le développement, ETH Zurich, Switzerl<strong>and</strong><br />

Abstract<br />

During the last few decades, a large extent of tropical rain forest has been cleared in the East of<br />

Madagascar through agricultural activities. Larger cohesive forest massifs are increasingly<br />

fragmented <strong>and</strong> forest fragments of various sizes remain in a l<strong>and</strong>scape mosaic dominated by<br />

agricultural patches. Until now, only little is known about how these fragments are perceived by<br />

the local population <strong>and</strong> what role they play in the local livelihood systems. We therefore tried<br />

to get a holistic underst<strong>and</strong>ing about the human-forest interface in this fragmented l<strong>and</strong>scape<br />

<strong>and</strong> based our methodology on the sustainable livelihood approach. The research has been<br />

conducted in four villages which differ in their distance to the cohesive forest massif <strong>and</strong><br />

therefore in their access to forest resources. We recognized that the perception of forest<br />

fragments <strong>and</strong> their importance for the local population changed with increasing distance to the<br />

forest massif. Not only they became more important to satisfy people’s daily needs, but they<br />

also had an increasing potential to cause conflicts between villagers. For a possible<br />

improvement of forest management designs, we should therefore take into account what role<br />

forest fragments play in local livelihood systems <strong>and</strong> how they can vary with changing access to<br />

forest resources.<br />

1 Introduction<br />

On a global level, the planet is continuously loosing its original tropical forests. Most tropical<br />

l<strong>and</strong>scapes not only suffer from deforestation but also from fragmentation, which often leads to<br />

decreasing vitality of remaining forest patches (Shvidenko et al. 2008). This is also the case in<br />

Madagascar, where forests are increasingly fragmented by agricultural activities such as slash<strong>and</strong>-burn<br />

cultivation (Harper et al. 2008). Nevertheless, these fragments are of increasing<br />

importance, not only for the diversity of mosaic l<strong>and</strong>scapes but also for the local inhabitants in<br />

the vicinity of these l<strong>and</strong>scapes (Pfund et al. 2006). Currently, in Madagascar forest resources<br />

are managed without separation between larger forests <strong>and</strong> fragments <strong>and</strong> without a broad<br />

consideration of local people’s livelihood strategies. We therefore aimed to explore the effective<br />

importance of forest fragments in local livelihood systems with our research. We combined<br />

quantitative <strong>and</strong> qualitative analyses about opportunities resulting from forest resources for<br />

people’s livelihood <strong>and</strong> the local perception about the importance of forested l<strong>and</strong>scapes. To<br />

explore the role of forested l<strong>and</strong>scapes, we divided them into two different categories; massif<br />

<strong>and</strong> fragments. This short paper will present two aspects of the role of forest l<strong>and</strong>scapes in local<br />

people’s life; the role of monetary income <strong>and</strong> the changing perception of forest fragments<br />

related to the distance to large non-fragmented forests.<br />

* Corresponding author. Phone: +4144 632 49 92<br />

Email address: zora.urech@env.ethz.ch<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Z.L. Urech & J.P. Sorg 2010. Taking into account local people’s livelihood systems for a better management of fragments<br />

629<br />

2 Methodology<br />

As a basis <strong>and</strong> general framework for our socioeconomic analyses we employed the Sustainable<br />

Livelihood Approach (SLA) (NADEL 2007). This approach allowed us to recognize the<br />

complexity of local people’s livelihood systems <strong>and</strong> strategies at individual, family <strong>and</strong><br />

community levels in relation to forest fragments. Diverse methods were used for data collection,<br />

based on the SLA. In general, our aim was to compare the perception of different l<strong>and</strong>scape<br />

types by the local population with concrete quantitative information, as for example about the<br />

income. We therefore worked with interviews <strong>and</strong> scoring exercises. All in all, we spent 11<br />

months in four villages during the two field periods of the project.<br />

2.1 Village selection<br />

For our research, we worked in four villages situated around the forest massif. The territories of<br />

the four villages have a different forest cover <strong>and</strong> differ especially in their distance to the forest<br />

massif, what correlates to the distance to markets (see Table 1). The villages Ambofampana <strong>and</strong><br />

Maromitety are situated near the massif in a remote area, whereas the villages Bevalaina <strong>and</strong><br />

Antsahabe are farther from the massif in a territory of lower forest cover <strong>and</strong> are less remote.<br />

The category near or far the massif is defined by the walking time to reach the border of the<br />

massif. The differences of the distance to the massif allowed us to analyse the influence of a<br />

changing l<strong>and</strong>scape on the human-forest interface.<br />

Table 1: The four selected villages<br />

Characteristics of village territories Ambofampana Maromitety Bevalaina Antsahabe<br />

Distance to forest massif<br />

[walking time in h] 0.25 0.5 2 3<br />

Category of distance to forest massif near near far far<br />

<strong>Forest</strong> cover [%] 86 75 43 21<br />

Market proximity [walking time in h] 6 8 2 1<br />

Number of households interviewed 25 22 32 31<br />

Number of groups for scoring exercises 6 4 6 6<br />

2.2 Interview methods<br />

We began with open-ended discussions with several households in each of our villages to get a<br />

general overview. We then conducted the first semi-structured household interviews to deepen<br />

particular topics. In these interviews we gathered more specific <strong>and</strong> also quantitative information<br />

about the importance, income, perceptions, products collected from <strong>and</strong> uses of forest<br />

fragments. To complete our data we held discussions with people who have specific knowledge<br />

or play a key role in the social context. These included older persons, loggers or traditional<br />

authorities.<br />

2.3 Scoring exercises<br />

To deepen the information on the perception of importance, we conducted scoring exercises<br />

with focus groups, separated by wealth levels <strong>and</strong> gender (Sheil <strong>and</strong> Liswanti 2006). To express<br />

their own perception of value, each group had to distribute 100 pebbles on 9 different l<strong>and</strong>scape<br />

types (Table 2) according to their importance. This had to be done 8 times for 8 different<br />

categories of goods <strong>and</strong> products (Table 3).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Z.L. Urech & J.P. Sorg 2010. Taking into account local people’s livelihood systems for a better management of fragments<br />

630<br />

Table 2: Categories of l<strong>and</strong>scape types<br />

L<strong>and</strong>scape types<br />

River<br />

Irrigated rice fields<br />

Tavy<br />

Safoka<br />

Marsh<br />

<strong>Forest</strong> massif<br />

Fragments<br />

Village garden<br />

Tanimboly<br />

Definition<br />

Water <strong>and</strong> riverside<br />

Irrigated, permanent rice fields<br />

Cultivation of rice <strong>and</strong> other products on slopes after slash-<strong>and</strong>-burn<br />

Secondary vegetation without cultivation<br />

Wet <strong>and</strong> periodically or permanent flooded ground<br />

Permanent natural tree cover connected to the forest massif.<br />

Permanent natural tree cover not connected to the forest massif.<br />

Trees <strong>and</strong> plants cultivated in the village around the houses<br />

Traditional agroforestry system with trees <strong>and</strong> annual crops<br />

Categories<br />

Food<br />

Medicine<br />

House construction<br />

Tools<br />

Fire wood<br />

Weaving<br />

Income<br />

Hunting<br />

Table 3: Categories of goods <strong>and</strong> products<br />

Definition<br />

Plants, products or animals which can be eaten<br />

Natural products used for medicine <strong>and</strong> health<br />

Materials to build houses<br />

Materials to build tools for agriculture, hunting<br />

Fuel<br />

Plants used for weaving products, such as mats, hats, baskets<br />

Products which can be sold<br />

Animals (lemurs, tenreks, fish etc.)<br />

3 Results<br />

3.1 The general importance of fragments<br />

<strong>Forest</strong>ed l<strong>and</strong>scapes are of indisputable high importance for the rural people living in the presented<br />

research area. But the local population perceive the massif <strong>and</strong> fragments as two types of<br />

l<strong>and</strong>scapes with different importance. As illustrated in Figure 1 the importance of fragments becomes<br />

significantly higher with increasing distance to the massif (Analysis of variance: F 3,18 =<br />

4.21, P = 0.02). The relatively high importance of fragments in the nearest village to the massif<br />

(Ambofampana) can be explained by the qualitative <strong>and</strong> quantitative still high availability of<br />

NTFPs <strong>and</strong> timber in these fragments. Because the population density is low, fragments are not<br />

much degraded <strong>and</strong> the difference of diversity between the massif <strong>and</strong> the fragments is not very<br />

high.<br />

Figure 1: The perceived importance of forested l<strong>and</strong>scapes, including all categories of goods <strong>and</strong> products<br />

(see Methodology, 2.3 Scoring exercises)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Z.L. Urech & J.P. Sorg 2010. Taking into account local people’s livelihood systems for a better management of fragments<br />

631<br />

Moreover, in our research area we found an important traditional rule concerning the right to<br />

convert forest fragments into agricultural l<strong>and</strong>. Only farmers being owners of the l<strong>and</strong><br />

surrounding the forest fragments have the right to clear these forests. This is a crucial rule as it<br />

is depending on the individual household whether a fragment will be cleared or not. Therefore<br />

we asked households which still have forest fragments situated next to their agricultural<br />

territories why they did not clear their fragment until this day. Figure 2 indicates the obvious<br />

changing interest for forest fragments with increasing distance to the massif. Households near<br />

the massif see fragments mainly as a soil reserve for future conversion into agricultural l<strong>and</strong>.<br />

Interesting is that this can not be a question of l<strong>and</strong> availability, as households near the massif<br />

have around twice as much available agricultural l<strong>and</strong> per households as families far from the<br />

massif do. Thus households far from the massif should be more interested in forest conversion,<br />

but they do not seem to be. Households far from the massif are more interested in the goods<br />

produced by forest fragments.<br />

Therefore our conclusion suggests that the perception of the importance of forest fragments<br />

changed with increasing scarcity of forest resources. On the one h<strong>and</strong>, people living far from the<br />

massif notice the growing concurrence on products from forested l<strong>and</strong>scapes. On the other h<strong>and</strong><br />

they have already been aware that forests are disappearing <strong>and</strong> are exhaustible. People near the<br />

massif generally think that forests are inexhaustible <strong>and</strong> they still have more than enough to<br />

satisfy their principal needs related to forests.<br />

Near massif<br />

Far massif<br />

Figure 2: Reasons for the preservation of forest fragments<br />

3.2 Changing importance of forests for income<br />

In this section we explore the importance of forested l<strong>and</strong>scapes more through the lense of<br />

monetary income resulting from forest products (see Table 1). Especially during periods of rice<br />

shortage, households are strongly depending on alternative income to buy food. Then, logging<br />

<strong>and</strong> timber transport become important sources of income. Farmers living near the forest massif<br />

still find precious wood in the massif, whereas farmers living far from the forest massif have to<br />

find wood in the surrounding fragments, where precious woods are already becoming scarce.<br />

Nevertheless, as shown in Figure 3, farmers living far from the massif have the higher income<br />

by timber activities than farmer living near the massif (Analysis of variance: F 3, 102 = 2.86, P =<br />

0.040).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Z.L. Urech & J.P. Sorg 2010. Taking into account local people’s livelihood systems for a better management of fragments<br />

632<br />

This is comprehensible as timber is tradable much better in villages far from the massif than<br />

near the massif. This existing trade far from the massif can be explained by the increasing<br />

scarcity of wood <strong>and</strong> the growing population farther away from the massif. Additionally, timber<br />

can be sold to families in other villages which can not walk this far to find wood. Another<br />

reason for the better trade far from the massif is the proximity to markets. If they live near the<br />

massif, farmers have to walk for 6-8 hours to reach the next market to sell their wood. This is<br />

almost impossible because farmers have to carry the wood on their shoulders.<br />

Contrary to the case of timber, the trade of non-timber-forest-products (NTFPs) is not<br />

significantly related to the massif proximity. On the one h<strong>and</strong>, NTFPs are still available in a<br />

higher amount <strong>and</strong> better quality in the massif than in the fragments (Fedele 2010), what allows<br />

a better trade for people living near the massif. On the other h<strong>and</strong>, NTFPs are easier to carry for<br />

long distances thus people living near the massif can walk 8 hours to the next market.<br />

Figure 3: Income generated by NTFP <strong>and</strong> timber from forested l<strong>and</strong>scapes<br />

From the results presented in Figure 3 we would assume that people far from the massif<br />

perceive forested l<strong>and</strong>scapes more important for income than people living near the massif. But<br />

this is not the case. People living near the massif perceive forested l<strong>and</strong>scapes (including both<br />

categories massif <strong>and</strong> fragments) not less important than people far from the massif. This can be<br />

seen in Figure 4, which shows the results of the distribution of 100 pebbles according to the<br />

perceived importance of different forested l<strong>and</strong>scapes for income (see 2.3 Scoring exercises).<br />

The results showed no significant difference for the relation between distance <strong>and</strong> importance of<br />

forested l<strong>and</strong>scapes. Obviously, the quantification of importance by income (Figure 3) does not<br />

reflect the actual perception of the local population (Figure 4). We asked the different groups<br />

for reasons explaining the given importance of forested l<strong>and</strong>scapes, even though the effective<br />

income from forest was not that high. The explanation was that the continuous availability of<br />

forest products was more crucial than the effective income. Products from forested l<strong>and</strong>scapes<br />

are always available <strong>and</strong>, although to a limited extent, always tradeable. This is a significant<br />

characteristic for crisis <strong>and</strong> periods of rice shortage. <strong>Forest</strong> products can not all be destroyed by<br />

cyclones, whereas crops are always in danger.<br />

Nevertheless, Figure 4 points out the increasing importance of fragments (Analysis of variance:<br />

F 3, 18 = 4.01, P = 0.024) <strong>and</strong> decreasing importance of the massif (Analysis of variance: F 3, 18<br />

= 3.23, P = 0.047) with growing distance between village <strong>and</strong> massif. These trends go into the<br />

same directions as already shown in Figure 1. While the results in Figure 1 are related to the<br />

importance of forested l<strong>and</strong>s for all different categories of goods <strong>and</strong> products, Figure 4 is only<br />

related to products that can be sold.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Z.L. Urech & J.P. Sorg 2010. Taking into account local people’s livelihood systems for a better management of fragments<br />

633<br />

Figure 4: The perceived importance of forested l<strong>and</strong>scapes for income, including only the category<br />

income (see 2.3 Scoring exercises)<br />

4 Discussion<br />

This short paper has strived to explain that by the local population, forest fragments <strong>and</strong> forest<br />

massifs are not perceived as the same. The importance of fragments is related to quantitative<br />

opportunities <strong>and</strong> local livelihood strategies, which must be understood in a holistic view of<br />

livelihood systems. We assume that if forest fragments are not taken into account in future<br />

forest management plans, this can lead to social conflicts between villagers. <strong>Forest</strong> management<br />

plans in Madagascar normally consider wider areas, including villages far <strong>and</strong> near the massif in<br />

the same management plans. Furthermore, there is no differentiation between fragments <strong>and</strong> the<br />

massif. But we observed that people make the difference between fragments <strong>and</strong> the massif.<br />

Additionally, farmers do perceive the importance of forest fragments differently with increasing<br />

distance to the massif. These different views <strong>and</strong> following strategies can lead to conflicts<br />

between villages. Therefore, we recommend that in regional planning of natural resources,<br />

different views <strong>and</strong> strategies must be identified <strong>and</strong> considered to counter possible sources of<br />

conflicts. Thus forested l<strong>and</strong>scapes should be explored in a holistic context of local people’s<br />

livelihood.<br />

References<br />

Fedele, G. 2010. Local management strategies <strong>and</strong> their influence on the distribution of<br />

P<strong>and</strong>anaceae in a forest mosaic l<strong>and</strong>scape at the East Coast of Madagascar. Masters<br />

Thesis. ETH, Zurich.<br />

Harper, G. J., M. K. Steininger, C. J. Tucker, D. Juhn, <strong>and</strong> F. Hawkins. 2008. Fifty years of<br />

deforestation <strong>and</strong> forest fragmentation in Madagascar. Environmental Conservation<br />

34:325-333.<br />

Nadel. 2007. Working with a Sustainable Livelihood Approach. NADEL, DEZA, Zürich,.<br />

Pfund, J.-L., T. O'Connor, P. Koponen, <strong>and</strong> J.-M. Boffa. 2006. Transdisciplinary research to<br />

promote biodiversity conservation <strong>and</strong> enhanced management of tropical l<strong>and</strong>scape<br />

mosaics. IUFRO L<strong>and</strong>scape Ecology Conference, Italy.<br />

Sheil, D. <strong>and</strong> N. Liswanti. 2006. Scoring the Importance of Tropical <strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> with<br />

Local People: Patterns <strong>and</strong> Insights. Environmental Management 38:126-136.<br />

Shvidenko, A., J. Sven Erik, <strong>and</strong> F. Brian. 2008. Deforestation. Encyclopedia of Ecology.<br />

Academic Press, Oxford Pages 853-859.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Measures of l<strong>and</strong>scape structure as ecological indicators <strong>and</strong><br />

tools for conservation


E.R. Diaz-Varela et al. 2010. Multiscale analysis of l<strong>and</strong> use heterogeneity <strong>and</strong> dissimilarity<br />

635<br />

Multiscale analysis of l<strong>and</strong> use heterogeneity <strong>and</strong> dissimilarity as a<br />

support for planning strategies<br />

Emilio R. Diaz-Varela 1* , Carlos J. Alvarez-López, Manuel F. Marey-Pérez & Pedro<br />

Álvarez-Álvarez 2<br />

1<br />

University of Santiago de Compostela, Spain<br />

2<br />

University of Oviedo, Spain<br />

Abstract<br />

The analysis of l<strong>and</strong>scape structure is commonly oriented to the comprehension of<br />

pattern:process relationships <strong>and</strong> the evolution of such across spatial <strong>and</strong> temporal scales.<br />

Nevertheless, an important – <strong>and</strong> often neglected – aspect of l<strong>and</strong>scape analysis is its potential<br />

to use the derived information as a support for planning, design <strong>and</strong> decision making in<br />

l<strong>and</strong>scape management processes. The results of quantitative analysis of the spatial variation of<br />

l<strong>and</strong>scape’s composition <strong>and</strong> configuration across scales can help to identify areas distinctive<br />

regarding their heterogeneity, thus indicating different needs for spatial planning.<br />

In this work, a methodology is explored for the analysis of forest l<strong>and</strong>scape pattern oriented to<br />

planning applications. In it, a series of sequential steps are taken in order to guide analysis into a<br />

diagnosis useful for decision making. It starts by defining heterogeneity areas at different scales<br />

by means of multiscale approach. It is expected that such an approach, based in a “multi-level”<br />

structural analysis, would be useful for the spatial design of planning strategies.<br />

Keywords: multiscale analysis, l<strong>and</strong>scape metrics, l<strong>and</strong> use heterogeneity, dissimilarity,<br />

l<strong>and</strong>scape ecological planning<br />

1. Introduction<br />

The analysis of the spatial arrangement of l<strong>and</strong>scapes is generally adopted as a major tool for<br />

the comprehension of processes <strong>and</strong> dynamics of the l<strong>and</strong>scapes, <strong>and</strong> the relationship among<br />

their constitutive elements. Beyond the utility of pattern analysis for the underst<strong>and</strong>ing of<br />

l<strong>and</strong>scapes, it can constitute a strong support of l<strong>and</strong>scape planning strategies. In this work, the<br />

utility of two different l<strong>and</strong>scape indices to analyse heterogeneity in order to derive conclusions<br />

for spatial planning is explored. To do so, simulated l<strong>and</strong>scapes are used to test the behaviour of<br />

the indices at multiple scales, <strong>and</strong> ideas of their application to real planning cases are derived.<br />

2. Methodology<br />

2.1. Artificial l<strong>and</strong>scapes<br />

Artificial l<strong>and</strong>scapes were generated using Simmap software (Saura, 2003), based on the<br />

Modified R<strong>and</strong>om Clusters Method (MRC) (Saura <strong>and</strong> Martínez-Millán, 2000). This method<br />

generates more realistic l<strong>and</strong>scapes than other neutral model software (Saura <strong>and</strong> Martínez-<br />

Millán, 2000), due to the patchy, irregular shape of the results, which made it adequate to<br />

simulate l<strong>and</strong> use planning scenarios. The programme allows to control the degree of patchiness<br />

by the variation in the generation parameters p, n, Ai, <strong>and</strong> both the Minimum Mapping Unit <strong>and</strong><br />

* Corresponding author. Tel.: +34 982 285 900 ext. 23629 - Fax: +34 982 285 926<br />

Email address: emilio.diaz@usc.es<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.R. Diaz-Varela et al. 2010. Multiscale analysis of l<strong>and</strong> use heterogeneity <strong>and</strong> dissimilarity<br />

636<br />

the extension of the map (Saura, 2003). We generated artificial l<strong>and</strong>scapes for three values of p:<br />

the value above which the size of the largest cluster in the map starts to vary (0.45), an<br />

intermediate value (0.50), <strong>and</strong> the value immediately below the percolation threshold (0.58).<br />

Values beyond the percolation threshold generated almost complete dominance of one of the<br />

l<strong>and</strong> cover classes in preliminary tests.<br />

We considered five l<strong>and</strong> cover classes, the proportion of each class following two types of<br />

distributions: “equiprobable” (each class 20%), <strong>and</strong> “non-equiprobable” (with classes<br />

distributed 50%; 30%; 10%; 5% <strong>and</strong> 5%). In the latter case, the aim was to resemble a real<br />

distribution of l<strong>and</strong> cover with one l<strong>and</strong> cover clearly, but not completely, dominant, being the<br />

respective modeled l<strong>and</strong> cover types Planted forest, Agriculture, Urban, Natural forest, <strong>and</strong><br />

Shrubl<strong>and</strong>. The “equiprobable” distribution is aimed to obtain results for an even distribution of<br />

results which can be used as a theoretical. In the “non-equiprobable” distributions, a more<br />

realistic l<strong>and</strong>scape response is expected. For each of these cases, four different MMUs were<br />

considered, for 1 pixel, 10 pixels, 50 pixels <strong>and</strong> 99 pixels, i.e. a total of 24 different l<strong>and</strong>scape<br />

cases. Each case was coded with a sequential number representing the generation parameters,<br />

e.g.: “N5m1-58a” for 5 classes, MMU=1; p= 0.58; “a” type of class proportion, being “a”<br />

equiprobable, <strong>and</strong> “b” non-equiprobable.<br />

2.3. Analysis of l<strong>and</strong>scape heterogeneity <strong>and</strong> dissimilarity<br />

A common approach for the analysis of l<strong>and</strong>scape heterogeneity is the application of l<strong>and</strong>scape<br />

pattern metrics to a categorical map (O’Neill et al., 1988; Botequilha-Leitao <strong>and</strong> Ahern, 2002;<br />

McGarigal et al., 2002; Botequilha-Leitao et al., 2006). Although l<strong>and</strong>scape metrics are<br />

extensively used, careful consideration is required as to which are the most adequate as regards<br />

the aims of the study <strong>and</strong> the spatial data available (Li <strong>and</strong> Wu, 2004; Corry <strong>and</strong> Nassauer,<br />

2005; Diaz-Varela et al., 2009b; Dramstad, 2009). One of the problems with application of<br />

l<strong>and</strong>scape metrics to an entire study area lies in that metrics do not reveal the spatial distribution<br />

of the variables studied, <strong>and</strong> their scale-dependence, making necessary a system for subdivision<br />

into intermediate scales. Previous studies on the subject made by the authors revealed the utility<br />

of moving-window approaches in calculation of l<strong>and</strong>scape indices (Botequilha-Leitao <strong>and</strong><br />

Díaz-Varela, 2009; Díaz-Varela et al., 2009a), namely those related to information theory, like<br />

the Shannon-Wiener index. The Shannon-Wiener index was developed as a measure of the<br />

information content in a code (Shannon <strong>and</strong> Weaver, 1949), <strong>and</strong> is calculated by the expression<br />

(1):<br />

SHDI<br />

m<br />

= −∑<br />

p i<br />

⋅ log p<br />

i=<br />

1<br />

i<br />

(1)<br />

Where p i is the proportion of the l<strong>and</strong>scape occupied by the class type i, <strong>and</strong> m the total number<br />

of classes.<br />

Dissimilarity was analysed by means of contrast metrics, namely the Contrast Weighted Edge<br />

Density (CWED). CWED is calculated following the expression (2) (McGarigal et al., 2002):<br />

CWED =<br />

m<br />

m<br />

∑∑<br />

( e ik<br />

dik<br />

)<br />

= i<br />

A<br />

(10000)<br />

(2)<br />

i= 1 k + 1<br />

Where e ik is the total length (m) of edge in l<strong>and</strong>scape between patch types (classes) i <strong>and</strong> k;<br />

includes l<strong>and</strong>scape boundary segments involving patch type i; d ik is the dissimilarity (edge<br />

contrast weight) between patch types i <strong>and</strong> k; <strong>and</strong> A, the total l<strong>and</strong>scape area (m 2 ). d ik was<br />

estimated by our subjective criteria over the dissimilarity among l<strong>and</strong> cover types, <strong>and</strong><br />

represented in the following table:<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.R. Diaz-Varela et al. 2010. Multiscale analysis of l<strong>and</strong> use heterogeneity <strong>and</strong> dissimilarity<br />

637<br />

Table 1: Contrast weights among modeled l<strong>and</strong> cover types<br />

L<strong>and</strong> cover<br />

Agriculture<br />

Planted<br />

forest<br />

Natural<br />

forest<br />

Urban<br />

Shrubl<strong>and</strong><br />

Agriculture 0.0 0.6 0.4 0.2 0.3<br />

Planted forest 0.6 0.0 0.8 0.7 0.2<br />

Natural forest 0.4 0.8 0.0 0.7 0.3<br />

Urban 0.2 0.7 0.7 0.0 0.1<br />

Shrubl<strong>and</strong> 0.3 0.2 0.3 0.1 0.0<br />

For its calculation, FRAGSTATS software (McGarigal et al., 2002) was used, <strong>and</strong> a circular<br />

window shape was chosen. Window radii of 10, 20, 30… 100 cells were applied in the<br />

calculations. When a window covers no data, a border effect is caused producing null areas<br />

proportional to the window radius. To avoid it, the original map dimension was fixed so as to<br />

obtain at least a result map of 600x600 cells. Further post-processing was carried out with<br />

ESRI’s ArcGIS 9.2 software.<br />

3. Results<br />

A total of 24 of simulated maps were obtained, showing different spatial patterns corresponding<br />

to the initial generation parameters. From them, the calculation of the indices using the moving<br />

window approach resulted in a total of 240 different maps, showing the scale behavior of spatial<br />

distribution of l<strong>and</strong>scape heterogeneity at different scales (see Figure 1 for examples on the<br />

different indices). In those maps generated by lower window sizes, homogeneous zones (i.e.<br />

lower index values) are shown corresponding with patches with extensions larger than the<br />

window size. For the same l<strong>and</strong>scapes, complex borders or fine-scaled l<strong>and</strong>scapes (i.e., areas<br />

with patches with extensions lower than the window size) correspond to peaks in the<br />

heterogeneity values. Maps generated for larger window sizes tend to average local effects.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.R. Diaz-Varela et al. 2010. Multiscale analysis of l<strong>and</strong> use heterogeneity <strong>and</strong> dissimilarity<br />

638<br />

Figure 1: 3D representation of heterogeneity l<strong>and</strong>scape N5m99-58b (see text for interpretation of code) at<br />

diverse scales for SHDI (up) <strong>and</strong> CWED (down). The original l<strong>and</strong>scape <strong>and</strong> the corresponding window<br />

size are illustrated for comparison..<br />

This general response can be summarized as lower mean values for the smaller window sizes,<br />

where also the greater variance is attained, followed by a trend to stabilization in values<br />

corresponding with a decrease in variance, as window sizes grow. Figure 2 shows how this<br />

general trend is reflected in SHDI depending in the generation parameters (i.e., arrangement of<br />

l<strong>and</strong>scape elements) for three different examples. Equiprobable l<strong>and</strong>scapes with low values for<br />

m <strong>and</strong> p show a swift trend to stabilization in index values, following an asymptotic trend (Fig.<br />

2, upper left). A similar trend towards stabilization, but showing least decrease in variance is<br />

shown for non-equiprobable l<strong>and</strong>scapes (Fig. 2, lower). Some situations can also show a<br />

smother trend towards estabilization, corresponding with more complex structures in the<br />

l<strong>and</strong>scapes, as patches diverging in size <strong>and</strong> different interspersion patterns (Fig.2, upper right).<br />

Figure 2: Different scale behaviour of l<strong>and</strong>scapes as shown by changes in SHDI values with the window<br />

size. Dots represent mean values, boxes st<strong>and</strong>ard deviation (SD) values, <strong>and</strong> whiskers 1.96*SD. A portion<br />

of the l<strong>and</strong>scape is illustrated, <strong>and</strong> the window sizes super-imposed for comparison. See text to interpret<br />

the generation parameters by the l<strong>and</strong>scape code.<br />

For the case of CWED, the response shown by the index is slightly different, due to the<br />

calculation parameters of the indices (Figure 3). The maximum value of SHDI is attained for a<br />

equiprobable distribution of l<strong>and</strong> cover proportions. Nevertheless, as CWED calculates edge<br />

density, the effect of growing window area is translated in decreasing values for st<strong>and</strong>ard<br />

deviation, <strong>and</strong> initially decreasing, then increasing values for the mean.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.R. Diaz-Varela et al. 2010. Multiscale analysis of l<strong>and</strong> use heterogeneity <strong>and</strong> dissimilarity<br />

639<br />

Figure 3:Scale behaviour of l<strong>and</strong>scape N5m99-58b as interpreted by CWED. Though stabilization of<br />

values towards the higher window values is evident, this is not clearly asymptotic with a maximum value<br />

as in the case of SHDI. Dots represent mean values, <strong>and</strong> boxes st<strong>and</strong>ard deviation (SD) values<br />

4. Consequences <strong>and</strong> applicability in l<strong>and</strong>scape planning<br />

The utility of the approach is derived from three different consequences of the analysis.<br />

First, the behavior of the indices with changes in scale allows the detection of self-similarity in<br />

l<strong>and</strong>scapes. In figure 2 (upper left), this would correspond with the stabilization of the trend of<br />

the mean values of SHDI. For such cases, the heterogeneity as detected by the index presents<br />

little variation with the scale, thus l<strong>and</strong>scapes can be considered as self-similar. This would<br />

allow to identify characteristic scales for l<strong>and</strong>scapes, <strong>and</strong> areas where planning actions should<br />

be independent from the scale of application. However, this case would be rarely verified in real<br />

scenarios, <strong>and</strong> trends are more similar to the graphic shown in figure 2 upper right. In those<br />

cases, variability as shown by st<strong>and</strong>ard deviation, <strong>and</strong> the smooth trend of the mean can be<br />

interpreted as several types of l<strong>and</strong>scape sharing the same area, which should be identified.<br />

Second, comparison between results of SHDI <strong>and</strong> CWED shows that l<strong>and</strong>scape heterogeneity<br />

analysis could not be confined to spatial composition <strong>and</strong> configuration. The use of dissimilarity<br />

values in the calculation of l<strong>and</strong>scape indices integrates a semantic perspective in the analysis of<br />

l<strong>and</strong>scape pattern, <strong>and</strong> allows for a qualitative approach. By using contrast weights among l<strong>and</strong><br />

cover types, we can identify possible conflicting areas, to which specific planning actions can be<br />

targeted.<br />

Figure 4:Multi-scale application of the heterogeneity analysis. In the upper part, sequential process for<br />

delimitation of heterogeneity trends based in dissimilarity among l<strong>and</strong> cover types. In the part below,<br />

micro-scale heterogeneity detected as high-dissimilarity sites inside low-dissimilarity areas. See text for<br />

details<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E.R. Diaz-Varela et al. 2010. Multiscale analysis of l<strong>and</strong> use heterogeneity <strong>and</strong> dissimilarity<br />

640<br />

Finally, the third consequence, related to the previous ones, derives from multi-scale effects in<br />

the composition, configuration, <strong>and</strong> semantic structure of the l<strong>and</strong>scape. Figure 4 shows how<br />

high-contrast <strong>and</strong> low-contrast areas can be identified in a l<strong>and</strong>scape, taking the map created<br />

with the window size from which the stabilization trend is established, <strong>and</strong> using the mean value<br />

of CWED as a reference. Nevertheless, for more detailed scales, possible points of conflict can<br />

be detected as high-contrast areas. Such differences in dissimilarity among l<strong>and</strong> cover can be<br />

addressed in planning by the adoption of different strategies corresponding with the planning<br />

spatial level. For instance, the general map can resemble a regional approach to heterogeneity<br />

detection, while the more detailed scale can refer local effects. This kind of analysis presents a<br />

powerful field of application when multi-scale strategies of planning are developed.<br />

References<br />

Botequilha-Leitao, A., Ahern, J., 2002. Applying l<strong>and</strong>scape ecological concepts <strong>and</strong> metrics in<br />

sustainable l<strong>and</strong>scape planning. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 59: 65-93.<br />

Botequilha-Leitao, A., Díaz-Varela, E., 2009. Paradigmas emergentes no Ordenamento do<br />

Território e da Paisagem, e no Urbanismo: Nova oportunidade para a conservaçao do<br />

recurso solo. 2º Encontro Anual da Sociedade Portuguesa de Conservaçao do Solo.<br />

Universidade do Algarve, Campus de Gambelas, 8-10th July, 2009.<br />

Botequilha-Leitao, A., Miller, J., Ahern, J., McGarigal, K., 2006. Measuring l<strong>and</strong>scapes : A<br />

planner’s h<strong>and</strong>book. Isl<strong>and</strong> Press, Washington.<br />

Corry, N.C., Nassauer, J.I., 2005. Limitations of using l<strong>and</strong>scape pattern indices to evaluate the<br />

ecological consequences of alternative plans <strong>and</strong> designs. L<strong>and</strong>scape <strong>and</strong> Urban<br />

Planning, 72: 265-280.<br />

Diaz-Varela, E., Álvarez-López, C.J., Marey-Pérez, M., 2009a. Multiscale delineation of<br />

l<strong>and</strong>scape planning units based on spatial variation of l<strong>and</strong>-use patterns in Galicia, NW<br />

Spain. L<strong>and</strong>scape <strong>and</strong> Ecological Engineering, 5: 1-10<br />

Diaz-Varela, E., Marey-Pérez, M., Rigueiro-Rodríguez ,A., Álvarez-Álvarez, P., 2009b.<br />

L<strong>and</strong>scape metrics for characterization of forest l<strong>and</strong>scapes in sustainable management<br />

framework: Potential application <strong>and</strong> prevention of misuse. Annals of <strong>Forest</strong> Science, 66:<br />

301.<br />

Dramstad, W.E., 2009: Spatial metrics - useful indicators for society or mainly fun tools for<br />

l<strong>and</strong>scape ecologists. Norsk Geografisk Tidsskrift - Norwegian Journal of Geography,<br />

63: 246 — 254<br />

Li, H., Wu, J., 2004. Use <strong>and</strong> misuse of l<strong>and</strong>scape indices. L<strong>and</strong>scape Ecology, 19: 389-399.<br />

McGarigal, K., Cushman, S.A., Neel, M.C., Ene, E., 2002. FRAGSTATS: Spatial Pattern<br />

Analysis Program for Categorical Maps. Available on internet, URL:<br />

www.umass.edu/l<strong>and</strong>eco/research/fragstats/fragstats.html [Accessed May 8, 2009]<br />

O’Neill, R.V., Krummel, J.R., Gardner, R.H., Sugihara, G., Jackson, B., DeAngelis, D.L.,<br />

Milne, B.T., Turner, M.G., Zygmunt, B., Christensen, S.W., Dale, V.H., Graham, R.L.,<br />

1988. Indices of l<strong>and</strong>scape pattern. L<strong>and</strong>scape Ecology,1: 153-162.<br />

Saura, S. <strong>and</strong> Martínez-Millán, J., 2000. L<strong>and</strong>scape patterns simulation with a modified r<strong>and</strong>om<br />

clusters method. L<strong>and</strong>scape Ecology, 15: 661-678.<br />

Saura, S., 2003. SIMMAP 2.0. L<strong>and</strong>scape categorical spatial patterns simulation software.<br />

User’s Manual. Available on the internet, URL: http://www.udl.es/usuaris/saura. [Accessed<br />

November, 21 2008].<br />

Shannon, C.E., Weaver, W., 1949. The mathematical theory of communication. University of<br />

Illinois Press, Urbana, Illinois.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.L. Hernández-Stefanoni et al. 2010. Effects of l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong> age on species richness <strong>and</strong> biomass<br />

641<br />

Effects of l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong> age on species richness <strong>and</strong><br />

biomass in a tropical dry forest<br />

J. Luis Hernández-Stefanoni * , Juan Manuel Dupuy, Fern<strong>and</strong>o Tun-Dzul & Filogonio<br />

May-Pat<br />

Centro de Investigación Científica de Yucatán A.C. Unidad de Recursos Naturales,<br />

Calle 43 # 130. Colonia Chuburná de Hidalgo. C.P. 97200, Mérida, Yucatán. México<br />

Abstract<br />

Tropical dry forests are the terrestrial ecosystem with the largest extent in the Yucatán Peninsula,<br />

but have been scarcely studied <strong>and</strong> are poorly represented under protected areas. This study<br />

aims to characterize relationships between structure of vegetation <strong>and</strong> l<strong>and</strong>scape structure <strong>and</strong><br />

habitat type (st<strong>and</strong> age) considering different spatial scales. Species richness <strong>and</strong> biomass were<br />

calculated from 276 sampling sites, while l<strong>and</strong> cover classes were obtained from multi-spectral<br />

satellite image classification using Spot 5 satellite imagery. Species richness <strong>and</strong> biomass were<br />

related to patch age, l<strong>and</strong>scape metrics of patch types (area, edge, shape, similarity <strong>and</strong> contrast)<br />

<strong>and</strong> principal coordinate of neighbor matrices (PCNM) variables using regression analysis.<br />

PCNM analysis was performed to interpret results in terms of spatial scales as well as to<br />

decompose variation into spatial, age <strong>and</strong> l<strong>and</strong>scape structure components. Results indicate the<br />

best l<strong>and</strong>scape configuration to promote biodiversity conservation <strong>and</strong> to support carbon<br />

sequestration.<br />

Keywords: Biomass; variation partitioning; Species diversity; Spatial scales; forest succession<br />

1. Introduction<br />

Tropical dry forests (TDF) are the most extensive l<strong>and</strong> cover type in the tropics. More<br />

than half of tropical dry forests occur in the Americas, <strong>and</strong> Mexico contains 38% of the TDF in<br />

the continent. However TDF are the most threatened ecosystem in the world as a consequence<br />

of human activities (Portillo-Quintero <strong>and</strong> Sanchez-Azofeifa, 2010). Consequently, information<br />

about the ecological drivers of community structure at various spatial scales is fundamental to<br />

fully underst<strong>and</strong> <strong>and</strong> design effective strategies for conservation <strong>and</strong> management.<br />

Compared to other ecosystems, ecological studies in TDF are relatively new. In addition,<br />

most studies on the effects of fragmentation <strong>and</strong> l<strong>and</strong>scape patterns on plant communities focus<br />

on particular patches <strong>and</strong> on local species richness (α-diversity), while few studies examine<br />

different patch types at the whole l<strong>and</strong>scape level <strong>and</strong> address effects on attributes of<br />

community structure such as abundance or biomass (Hill <strong>and</strong> Curran, 2003). In this study we<br />

assess woody species density <strong>and</strong> biomass, <strong>and</strong> their response to l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong><br />

age to address the following questions: Which l<strong>and</strong>scape configurations <strong>and</strong> stages of forest<br />

succession maximize biological diversity What is the relative importance of l<strong>and</strong>scape<br />

structure <strong>and</strong> st<strong>and</strong> age for carbon storage<br />

Most research examining the effects of l<strong>and</strong>scape structure on species diversity <strong>and</strong><br />

structure of vegetation has focused on a single spatial scale. Yet, the degree to which l<strong>and</strong>scape<br />

* Corresponding author: Email address: jl_stefanoni@cicy.mx<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.L. Hernández-Stefanoni et al. 2010. Effects of l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong> age on species richness <strong>and</strong> biomass<br />

642<br />

configuration affects the structure <strong>and</strong> composition of plant communities depends on processes<br />

occurring at different spatial scales (Cushman <strong>and</strong> McGarigal, 2004). In order to assess the<br />

relationship between community attributes <strong>and</strong> l<strong>and</strong>scape structure across different scales it is<br />

necessary to consider scale-dependent ecological processes (Bellier et al, 2007). Thus, the goal<br />

of this study was to relate plant species density <strong>and</strong> biomass with l<strong>and</strong>scape structure, st<strong>and</strong> age<br />

<strong>and</strong> spatial variation of ecological structure at different spatial scales.<br />

2. Methods<br />

2.1 Study Area<br />

The study was conducted in a l<strong>and</strong>scape mosaic of 22 x 16 km 2 in Yucatán, México.<br />

The climate is tropical, warm, with summer rain <strong>and</strong> a dry season from November to April.<br />

Mean annual temperature is ca 26ºC, <strong>and</strong> mean annual precipitation between 1000 <strong>and</strong> 1200<br />

mm, concentrated between June <strong>and</strong> October. The l<strong>and</strong>scape consists of Cenozoic limestone<br />

hills with moderate slope (10-25°) alternating with flat areas, <strong>and</strong> the elevation ranges from 60<br />

to 160 m.a.s.l (Flores <strong>and</strong> Espejel 1994). The l<strong>and</strong>scape is dominated by seasonally dry<br />

semideciduous tropical forests of different ages of ab<strong>and</strong>onment after traditional slash-<strong>and</strong>-burn<br />

agriculture (Figure 1).<br />

2.2 Remotely sensed data <strong>and</strong> imagery processing<br />

A Spot 5 satellite image acquired on January 2005 was geo-referenced to Universal<br />

Transverse Mercator Projection (WGS 84) <strong>and</strong> radiometrically corrected to minimize the effect<br />

of atmospheric scattering. A false color composite image was created from b<strong>and</strong>s 2 (red), 3<br />

(near infrared), <strong>and</strong> 4 (mid infrared). This composite image was used as a spatial reference<br />

framework for selecting suitable training sites. At least three training sites were selected for<br />

each of the following l<strong>and</strong>-cover types: 1) 3-8 y-old secondary forest; 2) 9-15 y-old secondary<br />

forest; 3) >15 y-old secondary forest on flat areas; 4) >15 y-old secondary forest on hills; 5)<br />

agricultural fields; 6) urban areas <strong>and</strong> roads. The training areas <strong>and</strong> the false color image were<br />

used to perform a st<strong>and</strong>ard supervised classification of the Spot 5 b<strong>and</strong>s using the maximum<br />

likelihood algorithm (Idrisi Kilimanjaro V14.1, 2004). The overall accuracy <strong>and</strong> Cohen’s Kappa<br />

statistic were used to assess the accuracy of the map (Campbell, 1987).<br />

2.3 Calculation of l<strong>and</strong>scape pattern metrics<br />

We selected the following metrics that have been found relevant in l<strong>and</strong>scape studies<br />

(Mazerolle & Villard, 1999): patch density (PD), edge density (ED), mean area weighted shape<br />

index (SHAPE_AM), mean area weighted proximity index (PROX_AM), mean area weighted<br />

Euclidean nearest neighbor distance (ENN_AM) <strong>and</strong> total edge contrast index (TECI; see<br />

McGarigal et al. 2002 for a description of each metric). To calculate the proximity index, a<br />

search radius of 10 pixels (300 m) was used, which coincides with empirically derived evidence<br />

about the average size of a patch type. The weighted edge contrast between vegetation cover<br />

classes, required to compute the total edge contrast index, was calculated as the inverse of the<br />

Morisita-Horn similarity index between each pair of vegetation cover classes.<br />

2.4 Species density <strong>and</strong> biomass data<br />

Field data were recorded from a hierarchical plant survey conducted during the rainy<br />

season of 2009. First, 23 l<strong>and</strong>scapes of 1 km 2 were selected encompassing the whole range of<br />

forest fragmentation; within each l<strong>and</strong>scape, 12 sample sites were located following a stratified<br />

r<strong>and</strong>om design considering each of the four secondary forest cover classes (276 sampling sites<br />

in total). Each sampling site consisted of two concentric circular plots: all woody plants > 5 cm<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.L. Hernández-Stefanoni et al. 2010. Effects of l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong> age on species richness <strong>and</strong> biomass<br />

643<br />

in DBH (diameter at breast (1.3 m) height), hereafter referred to as adults, were sampled in a<br />

200 m 2 plot; whereas 1-5 cm DBH woody plants, hereafter referred to as juveniles, were<br />

sampled in a nested 50 m 2 subplot. In each site, we identified all woody plants, <strong>and</strong> measured<br />

their diameter <strong>and</strong> height. The number of adults, juveniles <strong>and</strong> all woody plant species was<br />

computed. To calculate above-ground biomass, two equations were employed: one for<br />

individuals ≥ 10 cm in DBH (Cairns et al, 2003) <strong>and</strong> the other for individuals < 10 cm in DBH<br />

(Hughes et al, 2000).<br />

2.5 Data analysis<br />

Multiple regression <strong>and</strong> variation partitioning methods (Bocard et al, 2004) were used to<br />

quantify the effects of l<strong>and</strong>scape structure <strong>and</strong> spatial dependence on species density <strong>and</strong><br />

biomass. First, a model using l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong> age variables was fit to the response<br />

variables using multiple regression. Then, a multiple regression model using spatial variables<br />

derived from a principal coordinate of neighbor matrices (PCNM) analysis of plot locations was<br />

fit to the response variables. Finally, the two models were combined into an overall regression<br />

model <strong>and</strong> variation partitioning was performed to determine the relatively importance of<br />

l<strong>and</strong>scape structure variables, st<strong>and</strong> age, pure spatial dependence <strong>and</strong> shared variation on species<br />

density <strong>and</strong> biomass.<br />

PCNM analysis <strong>and</strong> multiple regressions were used to identify l<strong>and</strong>scape structure <strong>and</strong><br />

st<strong>and</strong> age variables related to response variables at different spatial scales (Bocard et al, 2004).<br />

First, we partitioned the spatial model for each response variable into several additive<br />

submodels, <strong>and</strong> run a variogram analysis of significant PCNM vectors to identify their scale <strong>and</strong><br />

assign them to one of three groups: very board scale (distances of 8001 to 10500 m), broad scale<br />

(distances of 2001 to 8000 m), <strong>and</strong> local scale (distances of 0 to 2000 m). Then, we calculated<br />

predicted values of species density <strong>and</strong> biomass corresponding to each spatial submodel. Finally,<br />

a multiple regression model using st<strong>and</strong> age <strong>and</strong> l<strong>and</strong>scape structure metrics was fit to the<br />

predicted response variables for each submodel (very broad, broad <strong>and</strong> local scale).<br />

3. Results<br />

The l<strong>and</strong> cover thematic map of the study area is shown in Figure 1. This l<strong>and</strong>scape<br />

covers a total area of 37 242 ha; 94.3% is covered by forest in any of the four vegetation classes,<br />

<strong>and</strong> only 5.7% is covered by agriculture, urban areas <strong>and</strong> roads. The overall accuracy calculated<br />

for the map was 75.6%, <strong>and</strong> the Kappa index was 0.7.<br />

Figure 1: Location <strong>and</strong> l<strong>and</strong> cover map of the study area obtained from a supervised classification.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.L. Hernández-Stefanoni et al. 2010. Effects of l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong> age on species richness <strong>and</strong> biomass<br />

644<br />

Multiple regression results indicated statistically significant relationships between<br />

species density <strong>and</strong> st<strong>and</strong> age, PD, ED, SHAPE_AM <strong>and</strong> TECI (Table 1). Relationships<br />

involving PD <strong>and</strong> SHAPE_AM were consistently negative, indicating that plant diversity<br />

decreases as the number of patches increases <strong>and</strong> as the shape of a patch type becomes more<br />

irregular. On the other h<strong>and</strong>, biomass was best explained by st<strong>and</strong> age, PROX_AM, ENN_AM<br />

<strong>and</strong> TECI (Table 1). Biomass in all groups consistently responded negatively to ENN_AM <strong>and</strong><br />

TECI <strong>and</strong> positively to PROX_AM, indicating that biomass decreases with distance between<br />

patch types of the same vegetation or as the perimeter of the focal patch type increases its<br />

contrast with other patches types. For all groups of plants, st<strong>and</strong> age was positively associated to<br />

species density <strong>and</strong> biomass, except for juvenile biomass (Table 1).<br />

Table 1: Regression st<strong>and</strong>ardized coefficients for predicting species density <strong>and</strong> biomass from st<strong>and</strong> age<br />

<strong>and</strong> patch type metrics<br />

DEPENDENT PREDICTOR ALL WOODY ADULT JUVENILE<br />

VARIABLE VARIABLE PLANTS PLANTS PLANTS<br />

Biomass (R 2 = 0.68) (R 2 = 0.68) (R 2 = 0.16)<br />

AGE 0.725* 0.734* -0.316*<br />

PD<br />

ED<br />

SHAPE_AM -0.181*<br />

PROX_AM 0.110* 0.110* 0.107***<br />

ENN_AM -0.095** -0.096**<br />

TECI -0.080** -0.081**<br />

Species density (R 2 = 0.26) (R 2 = 0.44) (R 2 = 0.09)<br />

AGE 0.444* 0.634* 0.254*<br />

PD -0.218** -0.128***<br />

ED 0.322* 0.248** 0.190**<br />

SHAPE_AM -0.363* -0.349* -0.260*<br />

PROX_AM<br />

ENN_AM<br />

TECI -0.175**<br />

Variables included in the model with * p


J.L. Hernández-Stefanoni et al. 2010. Effects of l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong> age on species richness <strong>and</strong> biomass<br />

645<br />

Figure 2: Partitioning of the variation of species density <strong>and</strong> biomass using l<strong>and</strong>scape structure variables<br />

(L<strong>and</strong>), st<strong>and</strong> age (Age) <strong>and</strong> PCNM variables (space) for different groups of plants<br />

Table 2: Regression st<strong>and</strong>ardized coefficients of predictor variables (age <strong>and</strong> l<strong>and</strong>scape structure) that<br />

explain significant components of spatial patterns of biomass <strong>and</strong> species density at different scales<br />

DEPENDENT PREDICTOR Very Very Very<br />

VARIABLE VARIABLE Broad Broad Local Broad Broad Local Broad Broad Local<br />

Biomass (R 2 = 0.04) (R 2 = 0.07) (R 2 = 0.09) (R 2 = 0.04) (R 2 = 0.07) (R 2 = 0.09) (R 2 = 0.04) (R 2 = 0.04) (R 2 = 0.02)<br />

AGE 0.192* 0.271* 0.192* 0.271* -0.187* -0.202*<br />

PD<br />

ED<br />

SHAPE_AM<br />

PROX_AM 0.112*** 0.112*** 0.124**<br />

ENN_AM -0.147** -0.254* -0.147** -0.254*<br />

TECI -0.105*** -0.105***<br />

Species density (R 2 = 0.14) (R 2 = 0.07) (R 2 = 0.05) (R 2 = 0.05) (R 2 = 0.10)<br />

AGE 0.266* 0.231* 0.154** 0.224* -0.256*<br />

PD -0.134**<br />

ED 0.413* 0.201** 0.332*<br />

SHAPE_AM -0.277* -0.208* -0.231*<br />

PROX_AM<br />

ENN_AM<br />

TECI<br />

Variables included in the model with * p


J.L. Hernández-Stefanoni et al. 2010. Effects of l<strong>and</strong>scape structure <strong>and</strong> st<strong>and</strong> age on species richness <strong>and</strong> biomass<br />

646<br />

structure <strong>and</strong> spatial dependence had a comparable or even stronger effect on species diversity<br />

than st<strong>and</strong> age. These results are consistent with recent findings in tropical forests showing that<br />

long-term monitoring data follow chronosequence patterns for basal area, but not for species<br />

density, indicating that st<strong>and</strong> age largely determines basal area –<strong>and</strong> biomass, whereas species<br />

density seems to be strongly affected by other factors operating at the local <strong>and</strong> l<strong>and</strong>scape level<br />

(Chazdon et al. 2007).<br />

Our results also show that, for juveniles, pure spatial dependence appears to be the most<br />

important predictor of both species density <strong>and</strong> biomass. This result could either imply strong<br />

dispersal/recruitment limitation, or be due to an environmental component that was not<br />

considered, such as topographic or soil variables (Jones et al, 2008).<br />

PCNM submodels allowed us to link the spatial distribution of species density <strong>and</strong><br />

biomass to st<strong>and</strong> age <strong>and</strong> l<strong>and</strong>scape structure at different spatial scales. At the very broad scale,<br />

st<strong>and</strong> age contributed most to biomass, <strong>and</strong> l<strong>and</strong>scape structure to species density. At the broad<br />

scale, st<strong>and</strong> age contributed most to both species density <strong>and</strong> biomass.<br />

Finally, we found a strong negative association between species density <strong>and</strong> PD <strong>and</strong><br />

SHAPE_AM, as well as a positive association between species density <strong>and</strong> edge density (ED).<br />

These results may suggest that moderate levels of disturbance may enhance species richness in a<br />

forest-dominated l<strong>and</strong>scape. However, high levels of disturbance resulting in a high degree of<br />

fragmentation can have a strong negative impact on species richness (Hill <strong>and</strong> Curran, 2003).<br />

References<br />

Bellier, E., Monestiez, P., Durbec, J.P. <strong>and</strong> C<strong>and</strong>au, J.N. 2007. Identifying spatial relationships<br />

at multiple scales: principal coordinate of neighbour matrices (PCNM) <strong>and</strong> geostatistical<br />

approches. Ecography, 30: 385-399.<br />

Borcard D, Legendre P, Avois-Jacquet C, Tuomisto H. 2004. Dissecting the spatial structure of<br />

ecological data at multiple scales. Ecology 85:1826–1832<br />

Cairns, M.A., Olmsted, I., Granados, J., Argaez, J., 2003. Composition <strong>and</strong> aboveground tree<br />

biomass of a dry semi-evergreen forest on Mexico’s Yucatan Peninsula. <strong>Forest</strong>. Ecology<br />

<strong>and</strong> Management . 186, 125–132.<br />

Campbell, J.B. 1987. Introduction to remote sensing. The Guilford Press. New York.<br />

Chazdon, R.L., Letcher, S.G., Breugel, M. van,Martínez-Ramos, M. Bongers, F. <strong>and</strong> Finegan, B.<br />

2007. Rates of change in tree communities of secondary Neotropical forests following major<br />

disturbances. Phil. Trans. R. Soc. B 362: 273-289.<br />

Cushman SA, McGarigal K (2004) Patterns in the speciesenvironment relationship depend on<br />

both scale <strong>and</strong> choice of response variables. Oikos 105:117–124.<br />

Flores, J.<strong>and</strong> Espejel, I. 1994. Tipos de vegetación de la Península de Yucatán. Etnoflora<br />

Yucatanense, Fascículo 3. México 135 p.<br />

Hill J.L., Curran P.J. 2003. Area, shape <strong>and</strong> isolation of tropical forest fragments: effects on tree<br />

species diversity <strong>and</strong> implications for conservation. Journal of Biogeography 30: 1391-<br />

1403.<br />

Hughes, R.F., J.B. Kauffman <strong>and</strong> V.J. Jaramillo-Luque. 1999. Biomass, carbon, <strong>and</strong> nutrient<br />

dynamics of secondary forests in a humid tropical region of México. Ecology 80:1892-<br />

1907.<br />

Jones, MM., Tuomisto, H., Borcard, D, Legendre, P., Clark, DB., <strong>and</strong> Olivas, PC. 2008.<br />

Explaining variation in tropical plant community composition: inXuence of<br />

environmental <strong>and</strong> spatial data quality. Oecologia 155:593–604.<br />

Mazerolle M.J., Villard M.A. 1999. Patch characteristics <strong>and</strong> l<strong>and</strong>scape context as predictor of<br />

species presence <strong>and</strong> abundance: A review. Ecoscience 6(1): 177-124.<br />

McGarigal K., Cushman S.A., Neel M.C., Ene E. 2002. FRAGSTATS: spatial pattern analysis<br />

for categorical maps. University of Massachusetts.<br />

Portillo-Quintero, C.A., Sánchez-Azofeifa, G.A. 2010. Extent <strong>and</strong> conservation of tropical dry<br />

forests in the Americas. Biological Conservation 143 (2010) 144–155.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Uuemaa et al. 2010. Spatial gradients of l<strong>and</strong>scape metrics as an indicator of human influence on l<strong>and</strong>scape<br />

647<br />

Spatial gradients of l<strong>and</strong>scape metrics as an indicator of human<br />

influence on l<strong>and</strong>scape<br />

Evelyn Uuemaa * , Ramon Reimets, Tõnu Oja, Arno Kanal & Ülo M<strong>and</strong>er<br />

Department of Geography, Institute of Ecology <strong>and</strong> Earth Sciences, University of Tartu,<br />

Estonia<br />

Abstract<br />

L<strong>and</strong>scape metrics have been widely used for mapping of l<strong>and</strong> cover/use change <strong>and</strong> analysing<br />

relationships between l<strong>and</strong>scape pattern <strong>and</strong> processes. Natural factors <strong>and</strong> human activity<br />

diversify <strong>and</strong> homogenize l<strong>and</strong>scape simultaneously. The aim of our study was to analyze the<br />

extent <strong>and</strong> magnitude of human influence on l<strong>and</strong>scape along l<strong>and</strong>scape gradients near main<br />

roads. We calculated several l<strong>and</strong>scape metrics for gradients along main roads of Tartu <strong>and</strong><br />

Tallinn on Estonian Basic Map (1:10 000) from different years (1997/1998 <strong>and</strong> 2004/2005). We<br />

compared the gradients of l<strong>and</strong>scape metrics to determine how the anthropogenic areas have<br />

exp<strong>and</strong>ed in 10 years <strong>and</strong> whether the housing <strong>and</strong> infrastructure tend to exp<strong>and</strong> more quickly<br />

on fertile soils. The results showed that the housing has mostly exp<strong>and</strong>ed on fields eliminating<br />

the possibility to use fertile soils for agriculture or foresting. The pressure on forest areas was<br />

not as intense as on agricultural areas. The gradients enable to compare the changes in l<strong>and</strong>scape<br />

structure in time <strong>and</strong> space at the same time.<br />

Keywords: spatial gradients, l<strong>and</strong>scape metrics, settlement structure<br />

1. Introduction<br />

<strong>Change</strong>s in l<strong>and</strong>scape are caused by natural <strong>and</strong> anthropogenic factors. Natural changes occur<br />

during a long time, mostly due to climate change. The anthropogenic impact on l<strong>and</strong>scapes is<br />

expressed by changes in l<strong>and</strong> use related primarily to changes in the settlement structure.<br />

Natural factors <strong>and</strong> human activity diversify <strong>and</strong> homogenize l<strong>and</strong>scape simultaneously.<br />

L<strong>and</strong>scape metrics that means all parameters that quantify the spatial pattern of l<strong>and</strong>scape from<br />

topographic measures (Vivoni et al. 2005) to proportions of l<strong>and</strong> use/cover, shape <strong>and</strong> area<br />

metrics (Palmer 2004) have been suggested for measuring l<strong>and</strong>scape diversity. These metrics<br />

enable to evaluate habitat suitability for animals <strong>and</strong> birds (Weiers et al. 2004), nutrient fluxes<br />

<strong>and</strong> losses (Uuemaa et al. 2005). Attempts to relate l<strong>and</strong>scape metrics to human perception of a<br />

l<strong>and</strong>scape have been made as well (Antrop <strong>and</strong> Van Eetvelde 2000). Both human activities <strong>and</strong><br />

those of other biota in l<strong>and</strong>scape may be generalised as l<strong>and</strong>scape consumption (Oja, Prede<br />

2004).<br />

Different software has been developed to calculate l<strong>and</strong>scape metrics like FRAGSTATS<br />

(McGarigal <strong>and</strong> Marks 1995) but there are no unambiguously interpretable indicators <strong>and</strong> there<br />

can be found numerous researches on the issue of the use/interpretation of l<strong>and</strong>scape metrics<br />

(Wu 2004, Uuemaa et al. 2005, Oja et al. 2005, Uuemaa et al. 2007).<br />

* Corresponding author. Tel.: +372 52 27 830 - Fax: +372 7375 825<br />

Email address: evelyn.uuemaa@ut.ee<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Uuemaa et al. 2010. Spatial gradients of l<strong>and</strong>scape metrics as an indicator of human influence on l<strong>and</strong>scape<br />

648<br />

L<strong>and</strong> use changes caused by urbanization can severely affect biodiversity, energy flows,<br />

biochemical cycles, <strong>and</strong> climatic conditions (Baker et al. 2001). A systematically effective<br />

approach to analyze the effects of urbanization on ecosystems is to study the changes of<br />

ecosystem patterns <strong>and</strong> processes along an urban-to-rural gradient (McDonnell et al. 1997)<br />

focusing on the recognition of unique urban texture from the l<strong>and</strong>scape types (Weng 2007).<br />

Spatiotemporal gradient analysis enables to determine how urban centres have been enlarged in<br />

space <strong>and</strong> time.<br />

The aim of the study was to analyse how the housing areas have exp<strong>and</strong>ed in 1997−2005 in<br />

Estonia near two Estonian largest cities – Tartu <strong>and</strong> Tallinn <strong>and</strong> how is this process influenced<br />

by major roads. Second aim was to determine whether the new housing areas tend to exp<strong>and</strong> on<br />

fertile soils or not.<br />

2. Methodology<br />

We chose two study areas around two largest cities of Estonia (Figure 1). The size of the study<br />

areas was limited by the availability of the Estonian Basic Map (1:10 000) for two different<br />

years set: 1997/1998 <strong>and</strong> 2004/2005. Estonian Soil Map (1:10 000) was used for soil data. Soils<br />

were reclassified according to their suitability for agriculture (Kõlli 1994). This classification<br />

takes into consideration the fertility of soils <strong>and</strong> how suitable they are for agricultural activities.<br />

For determining whether the new housing has exp<strong>and</strong>ed mainly on fertile soils or not, we<br />

distinguished new settlements by using centroids of the buildings <strong>and</strong> spatial join. Analysis was<br />

performed with vector data as buildings require high accuracy of the map.<br />

For analysis of the influence of the roads on housing structure <strong>and</strong> l<strong>and</strong>scape structure we<br />

calculated several l<strong>and</strong>scape metrics with moving window method in Fragstats 3.3 (McGarigal<br />

<strong>and</strong> Marks 1995). We used edge density (ED), patch density (PD), mean shape index<br />

(SHAPE_MN), Simpson’s diversity index (SIDI) on l<strong>and</strong>scape level <strong>and</strong> patch density on<br />

class level. On these l<strong>and</strong>scape metric maps we calculated gradients along main roads 100m,<br />

200m... 1900m, 2000m from the road for different years to see how the settlement <strong>and</strong><br />

l<strong>and</strong>scape structure has changed in time.<br />

Figure 1. Study areas near Tallinn <strong>and</strong> Tartu<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Uuemaa et al. 2010. Spatial gradients of l<strong>and</strong>scape metrics as an indicator of human influence on l<strong>and</strong>scape<br />

649<br />

3. Results <strong>and</strong> discussion<br />

The analysis showed that on the Tartu study area most of the new buildings were built on fields<br />

(45,5%) in 1997−2005 <strong>and</strong> 33,3% of the new buildings were built on the yards. On the Tallinn<br />

study area most of the buildings were built on the yards (51,4%) <strong>and</strong> only 23,9% on the fields.<br />

74,7% of the Tartu study area had soils with good suitability for agricultural activities <strong>and</strong><br />

81,6% of the new buildings by area were built on these soils. In Tallinn study area only 56,2%<br />

had soils with good suitability for agricultural activities <strong>and</strong> 79,8% of the buildings by area were<br />

built on these soils in 1997−2005. The results showed that most of the new buildings are located<br />

on fields with good suitability for agricultural activities <strong>and</strong> the housing tended to exp<strong>and</strong> near<br />

main roads (Figure 2 <strong>and</strong> Figure 3).<br />

Figure 2. New housing areas on the study area of Tallinn <strong>and</strong> gradients along main roads (Tallinn-Narva,<br />

Tallinn-Tartu, Tallinn-Haapsalu). Soils are classified according to their suitability for agricultural<br />

activities (Kõlli, 1994).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Uuemaa et al. 2010. Spatial gradients of l<strong>and</strong>scape metrics as an indicator of human influence on l<strong>and</strong>scape<br />

650<br />

Figure 3. New housing areas on the study area of Tartu <strong>and</strong> gradients along main roads (Tartu-Valga,<br />

Tartu-Võru). Soils are classified according to their suitability for agricultural activities (Kõlli, 1994).<br />

The analysis of gradients showed that all the l<strong>and</strong>scape metrics analyzed indicate very well<br />

housing density as edge density, patch density <strong>and</strong> Simpson’s diversity index had higher values<br />

with higher housing densities <strong>and</strong> in case of mean shape index it was vice versa. The distance<br />

from the road appeared to play more important role than distance from the city in existing<br />

patterns of housing (Figure 4). However it can be assumed that housing densities decrease as<br />

we move away from the city <strong>and</strong> from the road but these results not show very clear trends. Part<br />

of it was beacause small settlements near roads that give higher peaks of edge density, patch<br />

density <strong>and</strong> Simpson’s diversity index. Nevertheless the l<strong>and</strong>scape was more fragmented in<br />

300m − 1000m then in 0m − 300m <strong>and</strong> 1000m − 2000m. This shows that the housing tends to<br />

exp<strong>and</strong> not directly near the main road but a little bit farther. People prefer to get quickly out of<br />

the city by main road <strong>and</strong> then move up to 1km away from the main road.<br />

800<br />

700<br />

edge density (m/ha)<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

100m<br />

300m<br />

500m<br />

1000m<br />

1500m<br />

2000m<br />

0<br />

2000 4000 6000 8000 10000 12000 14000 16000<br />

distance from city (m)<br />

Figure 4. Gradients of edge density for Tartu-Valga road in Tartu study area in year 2005. 100m, 300m,<br />

500m, 1000m, 1500m <strong>and</strong> 2000m mark the distance of the gradient from the road.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Uuemaa et al. 2010. Spatial gradients of l<strong>and</strong>scape metrics as an indicator of human influence on l<strong>and</strong>scape<br />

651<br />

The temporal analysis showed that spatial fragmentation has increased significantly from 1997<br />

to 2005 near main roads. The fragmentation has increased more quicly from 0m −1000m of the<br />

road <strong>and</strong> up to 10km from the cities (Figure 5). In distance 2000m from the main road the<br />

l<strong>and</strong>scape structure has not changed significantly (Figure 6).<br />

800<br />

edge density (m/ha)<br />

700<br />

year 1997<br />

600<br />

year 2005<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

2000 4000 6000 8000 10000 12000 14000<br />

distance from the city (m)<br />

Figure 5. Gradients of edge density for Tartu-Valga road in Tartu study area for years 1997 <strong>and</strong> 2005 at<br />

300m from main road.<br />

800<br />

edge density (m/ha)<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

year 1997<br />

year 2005<br />

0<br />

2000 4000 6000 8000 10000 12000 14000<br />

distance from the city (m)<br />

Figure 6. Gradients of edge density for Tartu-Valga road in Tartu study area for years 1997 <strong>and</strong> 2005 at<br />

2000m from main road.<br />

These results enable to better underst<strong>and</strong> the importance of infrastucture in suburbanisation<br />

process <strong>and</strong> use this knowledge in planning process. Housing areas built on fields have already<br />

faced serious problems because of the ruined amelioration systems <strong>and</strong> these problems with loss<br />

of fertile soils may deepen unless local municipalities do not direct planning process more<br />

effectively.<br />

4. Conclusions<br />

The housing areas have exp<strong>and</strong>ed significantly during years 1997−2005 <strong>and</strong> the l<strong>and</strong>scape<br />

fragmentation has also therefore increased. Housing areas also ted to exp<strong>and</strong> on fertile soils that<br />

are very suitable for agricultural activities. Therefore the area of the fertile soils is decreasing<br />

due to suburbaisation in Estonia. The housing areas exp<strong>and</strong> more quicly within 10km radius of<br />

the cities <strong>and</strong> within 1000m radius of the main roads. These methods <strong>and</strong> results can be used in<br />

planning process for decision suport.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Uuemaa et al. 2010. Spatial gradients of l<strong>and</strong>scape metrics as an indicator of human influence on l<strong>and</strong>scape<br />

652<br />

Acknowledgements<br />

This study was supported by Estonian Science Foundation grant No.8040 <strong>and</strong> Target Funding<br />

Project No. 0180049s09 of the Ministry of Education <strong>and</strong> Science, Estonia<br />

References<br />

Antrop, M., Van Eetvelde, V., 2000. Holistic aspects of suburban l<strong>and</strong>scapes: visual image<br />

interpretation <strong>and</strong> l<strong>and</strong>scape metrics. L<strong>and</strong>scape And Urban Planning, 50 (1-3), 43-58.<br />

Baker L.A., Hope D., Xu Y., Edmonds J. <strong>and</strong> Lauver L., 2001. Nitrogen Balance for the Central<br />

Arizona-Phoenix (CAP) Ecosystem. Ecosystems, 4, 582-602.<br />

Kõlli, R., 1994. Suitablility classes of soils for agricultural activities. Tartu. In Estonian.<br />

McGarigal, K., Marks, B. 1995. FRAGSTATS: spatial pattern analysis program for quantifying<br />

l<strong>and</strong>scape structure. USDA For. Serv. Gen. Tech. Rep. PNW-351.<br />

McDonnell M.J., Pickett S., Groffman P. <strong>and</strong> Bohlen P., 1997. Ecosystem processes along an<br />

urban-to-rural gradient. Urban Ecosysems, 1, 21-36.<br />

Oja, T. <strong>and</strong> Prede, M., 2004. L<strong>and</strong>scape consumption in Otepää, Estonia. In: Palang, H.,<br />

Sooväli, H., Antrop, M., Setten, G., (Eds). European Rural <strong>L<strong>and</strong>scapes</strong>: Persistence <strong>and</strong><br />

<strong>Change</strong> in a <strong>Global</strong>ising Environment. Kluwer, pp. 99-112.<br />

Palmer, J. F., 2004. Using spatial metrics to predict scenic perception in a changing l<strong>and</strong>scape:<br />

Dennis, Massachusetts. L<strong>and</strong>scape And Urban Planning, 69 (2-3), 201-218.<br />

Uuemaa, E., Roosaare, J. <strong>and</strong> M<strong>and</strong>er, Ü., 2005. Scale dependence of l<strong>and</strong>scape metrics <strong>and</strong><br />

their indicatory value for nutrient <strong>and</strong> organic matter losses from catchments. Ecological<br />

Indicators 5(4), 350-369.<br />

Vivoni, E.R., Teles, V., Ivanov, V.Y., Bras, R.L. <strong>and</strong> Entekhabi, D., 2005. Embedding<br />

l<strong>and</strong>scape processes into triangulated terrain models. International Journal of<br />

Geographical Information Science, 19(4), 429-457.<br />

Weiers, S., Bock, M., Wissen, M. <strong>and</strong> Rossner, G., 2004. Mapping <strong>and</strong> indicator approaches for<br />

the assessment of habitats at different scales using remote sensing <strong>and</strong> GIS methods.<br />

L<strong>and</strong>scape Urban Planning, 67 (1-4), 43-65.<br />

Wu, J.G., 2004. Effects of changing scale on l<strong>and</strong>scape pattern analysis: scaling relations.<br />

L<strong>and</strong>scape Ecology, 19 (2), 125-138.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L<strong>and</strong>scape assessment tools for adaptive management of<br />

tropical forested l<strong>and</strong>scapes


E. Penot 2010. Socio-economic diagnosis of a small region using an economic farming system modeling tool (Olympe)<br />

654<br />

Socio-economic diagnosis of a small region using an economic farming<br />

system modeling tool (Olympe). An approach from household to<br />

l<strong>and</strong>scape scales to assist decision making processes for development<br />

projects supporting conservation agriculture in Madagascar<br />

Eric Penot<br />

CIRAD UMR Innovation/URSCA-SCRID, Madagascar<br />

Abstract<br />

Two agricultural development projects based on conservation agriculture <strong>and</strong><br />

agriculture/livestock integration are implemented in Madagascar with both a “watershed<br />

approach” <strong>and</strong> a “farming system approach”: the BV-lac project in the area of Lake Alaotra <strong>and</strong><br />

the BVPI-SEHP project in Vakinankaratra (Central highl<strong>and</strong>s) <strong>and</strong> South-East. A farming<br />

systems reference monitoring network (FSRMN) has been set up with two objectives: i) to help<br />

the project in decision making processes for choosing appropriate technologies that will be<br />

developed according to a farmer’s typology using prospective analysis, ii) to monitor the<br />

project’s economical impact in the short <strong>and</strong> medium term. The farming system modelling<br />

approach is based on a software developed by INRA-CIRAD-IAMM (“Olympe”, JM Attonaty,<br />

INRA), The approach is based on partnership (smallholder, farmers’ organizations, project<br />

operators <strong>and</strong> local administration), farming system analysis, <strong>and</strong> modelling for a Decision<br />

Support Systems (DSS) project orientation. Adoption of conservation agriculture (CA)<br />

represents both a real change of paradigm for local farmers <strong>and</strong> a real challenge for agriculture<br />

<strong>and</strong> natural resources sustainability.<br />

Keywords: Farming system modelling, DSS (decision support system), conservation agriculture,<br />

Madagascar.<br />

Introduction<br />

A model has two main roles: a figurative role in representing the system (how it functions) <strong>and</strong><br />

a demonstrative role (possibilities <strong>and</strong> strategies). Combining these two roles leads to an<br />

explanatory model whose function is to represent a specific phenomenon that derives from<br />

general phenomena (management, accounting, <strong>and</strong> so on) as a function of the local conditions<br />

that characterise the farming systems. To underst<strong>and</strong> farming systems as a “productive system”<br />

<strong>and</strong> the logic behind technical choices recalls the “systemic approach”, widely used in the<br />

classical farming systems approach. The approach described here is based on partnership,<br />

farming system analysis <strong>and</strong> modelling for a Decision Support Systems (DSS) for development<br />

projects. In the past, methods <strong>and</strong> instruments were developed to help individual farmers make<br />

decisions (Attonaty et al., 1999). Today, we are faced with an increasing number of problems in<br />

which the several different stakeholders involved have also different interests. The aim is not to<br />

find THE optimal solution as do models based on linear programming or game theory but to<br />

create models that lead to acceptable compromises between the different stakeholders.<br />

1 Method; rationale for using the software “Olympe” for Farming Systems<br />

Modelling (FSM)<br />

Detailed knowledge of local farming systems <strong>and</strong> farmers’ strategies in different contexts such<br />

as pioneer zones, rehabilitation areas or traditional tree-crop belts can contribute to building<br />

improved <strong>and</strong> better adapted solutions to help farmers make the right decision about their future<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Penot 2010. Socio-economic diagnosis of a small region using an economic farming system modeling tool (Olympe)<br />

655<br />

investments at the right time. In collaboration with INRA <strong>and</strong> IAMM, CIRAD developed a<br />

software called “ Olympe ” that enables the modelling of farming systems (Penot 2003).<br />

Olympe is an economic modelling tool to develop farming simulations in order to help<br />

individual decision-making at farm level <strong>and</strong> may be used for project decision making. There is<br />

also a module at regional level with farm groups that allow the assessment of various types of<br />

flows between groups (farmers’ organisation, villages ….).<br />

Farming systems modelling associated with a farm typology can therefore be used to help<br />

projects in testing scenarios with various types of technologies (<strong>and</strong> risks) in order to assess<br />

what is the right technology for the right farmer at the right time according to farmers’<br />

strategies. Then, it aims to provide guidelines for agricultural <strong>and</strong> development policies for<br />

institutions <strong>and</strong>/or donors. Olympe can be used in a variety of situations <strong>and</strong> with different<br />

methodological approaches: comparison of cropping systems, the economics of farming<br />

systems <strong>and</strong> resource management (“farm management counselling” * ), prospective analysis,<br />

regional approach, <strong>and</strong> even for “role game.”<br />

Olympe simulator has been developed by J-M Attonaty (INRA Grignon, France) <strong>and</strong> associated<br />

partners from CIRAD <strong>and</strong> IAMM. It builds simulations by series of 10 years for one or more<br />

stakeholders, provides results <strong>and</strong> summarizes the results as a function of the needs of each<br />

stakeholder (Figure 1). Olympe is based on the systemic analysis of farming systems. The<br />

overall objectives of using Olympe are the following: i) to identify smallholders’ constraints<br />

<strong>and</strong> opportunities in a rapidly changing environment in preparation for the adoption of new<br />

cropping systems or any other organisational innovation <strong>and</strong> to underst<strong>and</strong> farmers’ strategies<br />

<strong>and</strong> their capacity for innovation., ii) to assess their ability to adapt to changing economic<br />

conditions, price crises <strong>and</strong> technological change, iii) to provide a tool to underst<strong>and</strong> the<br />

farmers’ decision-making process <strong>and</strong> to put information about farming systems in the social<br />

<strong>and</strong> economic context (through a regional approach), <strong>and</strong> iv) to undertake prospective analysis<br />

<strong>and</strong> build scenarios based on climatic risks <strong>and</strong> fluctuating commodity prices.<br />

It is also possible to calculate impact at the regional scale on various groups of farms (as a<br />

function of a given typology). Building scenarios trough prospective analysis allows to test the<br />

robustness of any decision or technical choice. Data analysis obtained with Olympe should be<br />

discussed with farmers in partnership in order to validate scenarios <strong>and</strong> guarantee a high degree<br />

of representativeness <strong>and</strong> accuracy. For instance, a network of selected representative farms can<br />

be monitored for several years to diagnose constraints <strong>and</strong> opportunities <strong>and</strong> to measure the<br />

impact of technical change. One of the main outputs of such an approach is the assessment of<br />

the impact of technical alternatives or choices from an economic <strong>and</strong> environmental point of<br />

view.<br />

Fig. 1: An iterative analysis of the problem.<br />

Data<br />

Scenario<br />

Model<br />

Results<br />

Analysis<br />

PROBLEM<br />

VIEW POINT<br />

Acceptable<br />

Compromise<br />

Possible<br />

Solutions<br />

* “Conseil de gestion” in French.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Penot 2010. Socio-economic diagnosis of a small region using an economic farming system modeling tool (Olympe)<br />

656<br />

2 Diversification <strong>and</strong> CA (Conservation Agriculture) as alternatives for<br />

sustainable development<br />

The sustainability of agriculture is becoming a major concern. Ecological <strong>and</strong> agricultural<br />

sustainability are linked trough degraded environment, fragile soils, fertility, biodiversity, <strong>and</strong><br />

the protection of watersheds. Crop diversification <strong>and</strong> rapid technical change characterise the<br />

evolution of existing farming systems. It is important therefore to analyze <strong>and</strong> underst<strong>and</strong> the<br />

key elements of the history of innovation processes so as to be in a position to make viable<br />

recommendations for development. Among other technologies, CA techniques are based on 3<br />

main items: no tillage, associated permanent covercrops <strong>and</strong> crop rotation. CA triggers a real<br />

change of paradigm for local farmers. Besides those constraints, CA techniques, though yields<br />

might not be significantly above that of tillage systems, provide a more sustainable production<br />

pattern through the climatic buffer effect of mulching <strong>and</strong> cover-crops.<br />

The notion of “economic sustainability” places emphasis on the profitability of specific<br />

technical choices such as analysis of margins, generation of income, return to labour <strong>and</strong> capital<br />

as a function of a specific activity, analysis of constraints <strong>and</strong> opportunities, etc...From the<br />

point of view of farming systems, both at the regional scale, <strong>and</strong> at the level of the<br />

“community” where there are serious constraints in l<strong>and</strong> availability, <strong>and</strong> in access to capital<br />

<strong>and</strong> information. Analysis of farming systems <strong>and</strong> knowledge about smallholders’ strategies in<br />

the different contexts are key factors that should be taken into account.<br />

Negotiations between stakeholders <strong>and</strong> better knowledge of the relations between the State <strong>and</strong><br />

farmers are essential if we are to improve the effectiveness of future projects <strong>and</strong> development<br />

actions. The main objective of topic-oriented research centred on the analysis of decisionmaking<br />

processes at different levels (farms, community, projects, <strong>and</strong> regional or national<br />

policy makers) would thus be to provide socio-economic information to policy makers to<br />

improve decision-making processes in agricultural development. The processes of innovation<br />

(farmers) <strong>and</strong> of decision-making (both farmers <strong>and</strong> developers) are key research topics in<br />

sustainable development. And the analysis of farming systems, the characterisation of agrarian<br />

systems <strong>and</strong> the identification of stakeholders’ strategies are key components to a better<br />

underst<strong>and</strong>ing of these issues. The factors that determine change in a sustainable development<br />

perspective need to be related to each specific context. Important issues such as the effect of<br />

decentralisation, globalisation <strong>and</strong> its effects on prices, as well as on local economies <strong>and</strong><br />

public policies, environmental topics (biodiversity, sustainability) are impossible to circumvent.<br />

One expected output would be the clear identification of the conditions required to ensure<br />

future projects are viable at the decision-making level. Farming system modelling through a<br />

farming system reference monitoring network provides a tool for technical choices made by<br />

decision makers with respect to agricultural policy.<br />

The main aim of this paper is to describe a possible global approach using a modelling tool<br />

which includes the identification of knowledge gaps <strong>and</strong> opportunities to promote actions <strong>and</strong><br />

projects or the implementation of policies that respect the need for sustainable development, as<br />

well as those of local stakeholders, developers <strong>and</strong> researchers. The historical dimension is very<br />

significant in this type of analysis even if economic commodity cycles can be very rapid. So<br />

far, rebuilding the past with a modelling tool <strong>and</strong> creating new evolution scenarios through<br />

prospective analysis can be linked to improve the efficiency of development-oriented research.<br />

The impact of technical change should take into account the effect of sustainability on both<br />

farmers‘ livelihood <strong>and</strong> on the environment. Success in diversification strategies requires a<br />

certain number of conditions: access to capital or credit, technical options (innovations), access<br />

to information, markets, <strong>and</strong> to farmers’ organisations in order to improve marketing, <strong>and</strong> so<br />

on.<br />

From farmers to developers<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Penot 2010. Socio-economic diagnosis of a small region using an economic farming system modeling tool (Olympe)<br />

657<br />

The use of Olympe enables a comprehensive underst<strong>and</strong>ing of how a given farming system<br />

functions <strong>and</strong> provides as well a tool to model prospective technical choices, price scenarios,<br />

<strong>and</strong> even ecological scenarios to test the robustness of technical choices. These tools can be<br />

used at different scales: that of the local community or that of regional, national or international<br />

scale, depending on the stakeholders <strong>and</strong> on the commodity involved. Emphasis should be on<br />

the farmers <strong>and</strong> on the other people directly involved in the farmers’ environment, including the<br />

government (development policies at the national level). Participatory <strong>and</strong> partnership<br />

approach, Action–Research (RD) are the main methodologies used in the approach.<br />

A prospective tool to assess the resilience of systems in the face of risk<br />

In this case the focus is on providing decision-making aid to administrators, projects, <strong>and</strong><br />

decision makers as well as to farmers themselves. Analysis of climatic events or the impact of<br />

price volatility, or any other economic risk allows the definition of scenarios where the<br />

resilience of a given farming system can be quantified. Care needs to be taken into account for<br />

the possible or induced perverse effects of “playing with scenarios” whose only validity is how<br />

representative they are. Olympe can also be used to reveal such induced or perverse effects. A<br />

typical example is that of the introduction of drop irrigation to save groundwater that eventually<br />

leads to over-consumption of water. The “revealing character” of FSM leads to enhanced<br />

sensitivity by stakeholders to problems that are not initially obvious. In this case, its use is very<br />

close to that of role game. Farming systems modelling can be used as a prospective tool to build<br />

scenarios about potential farm pathways, <strong>and</strong> to define agricultural policies, recommendations,<br />

to test the viability of recommendations as a function of local constraints, to assess different<br />

impacts, <strong>and</strong> the matching of policies to the real situation faced by the farmers. Risks analysis is<br />

a key component in this approach. Farming system modelling through a FSMN enables to<br />

effectively assess at the farm level risks <strong>and</strong> expected outputs from a choice.<br />

3 The References Farming System Monitoring Network (RFSMN): a<br />

comprehension tool of farmers’ strategies <strong>and</strong> follow-up evaluation.<br />

A References Farming System Monitoring Network (RFSMN) is a set of representative farms<br />

that show various agricultural situations dependent on soil/climatic units as well as socioeconomical<br />

situations, resulting from a typology. Farms are surveyed in-depth then followed<br />

<strong>and</strong> updated every year in order to measure i) the impact of the projects’ implementations, ii)<br />

the development policies in progress, iii) the resulting innovations’ processes (Penot, 2008).<br />

The objective through a follow-up is to measure the impact, the evaluation, the prospective<br />

analysis <strong>and</strong> decision-making process inside projects (choice of technologies to be promoted<br />

<strong>and</strong> level of intensification according to farm types for example…). A prospective analysis<br />

allows the comparison between potential scenarios <strong>and</strong> reality. The final objective is to allow<br />

development operators in contract with projects to measure impacts <strong>and</strong> re–orientate rapidly<br />

their actions.<br />

Parallel to the RFSMN, the project sets up procedures of plot <strong>and</strong> farms levels data acquisition<br />

whose objective is to obtain detailed <strong>and</strong> precise data allowing simulation <strong>and</strong> further<br />

prospective analysis,. A “plot database” common to all contracted operators allows the<br />

identification of cropping pattern, with data effectively observed in the fields that will feed the<br />

simulation. With the adoption of “farming system level approach”, rather than the traditional<br />

“plot level”, the project sets up “farming books”, on a voluntary basis in order to record farm<br />

evolution, description of cropping systems <strong>and</strong> main simple economic factors <strong>and</strong> analysis<br />

(gross <strong>and</strong> net margin, return to labour) <strong>and</strong> to observe tendencies <strong>and</strong> farms’ trajectories.<br />

Identification of a regional operational typology:<br />

The criteria of discrimination for the farm typology are the following: i) access to various types<br />

of soil with referring cropping systems (irrigated rice plantation, poor water management rice<br />

fields, upl<strong>and</strong> crops on “tanety”), ii) rice self-sufficiency <strong>and</strong> farm size, iii) level of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Penot 2010. Socio-economic diagnosis of a small region using an economic farming system modeling tool (Olympe)<br />

658<br />

intensification <strong>and</strong> use of inputs <strong>and</strong> production target (subsistence farming, sale…), iv) offfarm<br />

activities <strong>and</strong> diversification (agricultural productions <strong>and</strong> non agricultural activities, v)<br />

type of labour <strong>and</strong> material (manual, animal traction, motorization or combined traction) <strong>and</strong><br />

type <strong>and</strong> use of labour (familial <strong>and</strong> external).157 farms were surveyed in 2007/2008. For each<br />

identified type, four farms were modelled with the Olympe software, in 2007/2008, <strong>and</strong> were<br />

supplemented by a series of additional farms essential for a good follow-up/evaluation. The<br />

final network was composed of 48 farms. The databases of local operators (AVSF, BRL, SD-<br />

Mad…) provide reliable indicators on farmers’ technical plot pathways which are monitored by<br />

the project so as to build average st<strong>and</strong>ard cropping patterns. We need at least a minimum of 10<br />

plots with a homogeneous average of production (Coefficient of variation lower than 30%). The<br />

most complete database (from BRL/Madagascar), integrates 2800 plots. A complete review of<br />

the main results of these databases led to the identification of more than 120 cropping patterns<br />

that took into account: varieties, plot position on the transect <strong>and</strong> practices.<br />

Construction of st<strong>and</strong>ard cropping patterns according to the system of dichotomic keys.<br />

The use of simple dichotomic keys for selecting the right technologies apparently most adapted<br />

to local plot conditions (soils, climax, etc…) are currently used. Modelling “step by step” with<br />

Olympe is done in the form of a prospective analysis by testing scenarios differentiated<br />

according to the farming <strong>and</strong> socio-economic situations. The definition of the dichotomic keys<br />

remains a big step in the process of choosing technologies promoted by projects. We currently<br />

use several modalities to identify the adapted cropping pattern to be recommended: i) the use of<br />

local plot databases as presented above. ii) the use of the official recommendations from<br />

GSDM, synthesized in tables of description of cropping patterns from the CA h<strong>and</strong>book (O<br />

Husson et al., 2009) with generic dichotomic keys, iii) The use in the long term (2010) of a<br />

software tool specifically developed for selection of cropping patterns according to morphopedological<br />

constraints, “PRACT” developed by K Naudin (CIRAD/URD SCRID) in 2010.<br />

Indicators of management <strong>and</strong> measurement of risk<br />

The software enables the creation of scenarios based on various types of adoption <strong>and</strong><br />

modification of technical patterns (cropping or livestock), more or less intensive. Then, the<br />

objective is to test the robustness of technical choices, <strong>and</strong> then the impact on production<br />

systems caused by climatic risks (cyclones, output lower due to the attack on a plant’s health,<br />

excess or lack of water, etc…) or economic (impact of the volatility of the farm prices <strong>and</strong> the<br />

inputs). Indicators (st<strong>and</strong>ard formula Excel type) allow to calculate ratios <strong>and</strong> traditional<br />

economic variables of management. The identification of simple ratios <strong>and</strong> the consequent<br />

analysis of the financial farm situation after a technical choice, a real or simulated one, largely<br />

facilitated the appropriation by operators <strong>and</strong> led to a better integration of their<br />

recommendations, while taking into account the concepts of risk for the farmer. Such an<br />

approach allows operators to better include <strong>and</strong> underst<strong>and</strong> farmers’ strategies in production<br />

factors allowance <strong>and</strong> finally in the farmers’ priorities of resource allocation according to their<br />

knowledge, their own experimentation, their potential opportunities <strong>and</strong> their current situation.<br />

Risks lead to shocks <strong>and</strong> disturbances. Impact strength can be regarded as the capacity of a<br />

system to overcome disturbances while maintaining its vital functions, its structure <strong>and</strong> its<br />

capacities of control. It is thus important for the capacity of a system to be able to resist by<br />

maintaining the essence of its structure <strong>and</strong> “modus oper<strong>and</strong>i” while including the possibility of<br />

any change. It is based on the conditions which maintain an initial balance though potentially<br />

unstable which can lead to another balance. One can measure it by the magnitude or the level of<br />

disturbances a system can resist or absorb until the rupture or the change of that system’s<br />

structure. The robustness can then be interpreted like a particular impact strength according to a<br />

definition close to that used in statistics. Risks are assessed through the use of the “hazard<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Penot 2010. Socio-economic diagnosis of a small region using an economic farming system modeling tool (Olympe)<br />

659<br />

module” in Olympe which enables the creation of scenarios with any changes in inputs/output<br />

prices as well as production <strong>and</strong> yield.<br />

Conclusion<br />

3 RFSMN’s are currently been set up (lake Alaotra <strong>and</strong> Vakinankaratra). Farming system<br />

analysis, training <strong>and</strong> modelling with a simple tool (Olympe), linked with the use of existing<br />

plot databases managed by operators contributed largely to the effective development of a real<br />

“farming system approach” in these projects. Training <strong>and</strong> the use of the tool lead to a real<br />

pedagogic impact on various operators. Extentionists <strong>and</strong> staff managers start to adapt their<br />

recommendations <strong>and</strong> feel more empowered to take responsibility in their extension activities.<br />

Processes of innovations are better recorded <strong>and</strong> integrated into the analysis. The farming<br />

system approach allow as well to better consider other level of action such as farmers’<br />

organisations for services (required for CA adoption such as information, credit, access to<br />

inputs or marketing…) or the regional level (watershed, village area, community territory…).<br />

The case of CA is a strong example that has also largely contributed to the adaptation of its<br />

very particular <strong>and</strong> specific agricultural systems towards a stronger more encompassing<br />

adaptation (simplification, adaptation, medium/low intensification, increase of the possible<br />

range of the techniques functioning for the local specifics). The idea for support of the different<br />

services for agriculture (dispersion, credit, supply, commercialisation) were changed <strong>and</strong> their<br />

importance finally accepted by the operators whose initial goals were simple <strong>and</strong> concise: to<br />

have the maximum of parcels improved without regard for the type of exploitation. The<br />

installation of tools has therefore vastly contributed to the strengthening of the approach itself<br />

<strong>and</strong> its usage <strong>and</strong> acquisition by the development operators in a type of “learning by doing”<br />

training approach. A key element was equally the participation of the genuine partners since the<br />

beginning of the operation in July 2006 at Lac Alaotra. The concept, the approach, the<br />

donations, <strong>and</strong> the results were all explored, analyzed, <strong>and</strong> validated by the operators which in<br />

turn strengthen their will to underst<strong>and</strong>, master, <strong>and</strong> use the tools presented in this text.<br />

There still is a need to implement a value analysis in the near future of the results obtained from<br />

the more interesting scenarios compared to what effectively happened in the last 5 years.<br />

Bibiography<br />

Attonaty et al., 1999 J.-M. Attonaty, M.-H. Chatelin, <strong>and</strong> F. Garcia, "Interactive simulation<br />

modelling in farm decision-making," Computers <strong>and</strong> Electronics in Agriculture, vol.<br />

22..pp. 157-170.<br />

Husson O.; Charpentier H. ; Razanamparany C. Moussa N.; Razafintsalama H.; Michellon R.;<br />

Naudin K.; Rakotondramanana; Séguy L. . 2007. Les systèmes a proposer en priorité<br />

dans les différents milieux de Madagascar. Manuel pratique du semis direct à<br />

Madagascar, vol II. Antananarivo, Madagascar. GSDM. 165 p.<br />

Penot E. 2008. Document de travail du PROJET BV-LAC N° 4 : Mise en place du réseau de<br />

fermes de références avec les opérateurs du projet.. Projet BV-lac/AFD. 35 pp.<br />

Penot E, Deheuvels O, Attonaty JM, Le Grusse Ph. 2003. Modélisation des exploitation<br />

agricoles : les mutiples usages du logiciel Olympe., Montpellier, CIRAD.<br />

Communication au séminaire modélisation avec Olympe, Montpellier, 2003, 20 pp.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Network theory to conserve <strong>and</strong> reconnect forested<br />

l<strong>and</strong>scapes


S. Decout et al. 2010. Connectivity loss in human dominated l<strong>and</strong>scape<br />

661<br />

Connectivity loss in human dominated l<strong>and</strong>scape: operational tools for<br />

the identification of suitable habitat patches <strong>and</strong> corridors on<br />

amphibian’s population<br />

Samuel Decout 1 , Stéphanie Manel 2,3 , Claude Miaud 4 & S<strong>and</strong>ra Luque 1<br />

1 Cemagref, Unité de Recherches Ecosystèmes Montagnards, 2 Rue de la Papeterie,<br />

38402 Saint-Martin d'Hères, France<br />

2 Laboratoire d’Ecologie Alpine, Equipe Génomique des Populations et Biodiversité, BP<br />

53, 2233 Rue de la Piscine, 38041 Grenoble Cedex 9, France<br />

3 Laboratoire Population Environnement Développement, Equipe Ville Environnement<br />

Développement, Université Aix-Marseille 1, 3 place Victor Hugo, 13331 Marseille<br />

cedex 03, France<br />

4 Laboratoire d’Ecologie Alpine, Equipe Génomique des Populations et Biodiversité,<br />

Université de Savoie, Bâtiment Belledonne, 73376 Le Bourget du Lac, France<br />

Abstract<br />

L<strong>and</strong>scape connectivity is a key issue for biodiversity conservation. Many species have to<br />

refrain to move between scattered resources patches. This is particularly the case for the<br />

common frog, a widespread amphibian migrating between forest <strong>and</strong> aquatic habitats for<br />

breeding. Face to the growing need for maintaining connectivity between amphibians’ habitat<br />

patches, the aim of this study is to provide a method based on habitat suitability modelling <strong>and</strong><br />

graph theory to explore <strong>and</strong> analyze ecological networks. We first used the maximum entropy<br />

modelling with environmental variables based on forest patches distribution to predict habitat<br />

patches distribution. Then, with considerations about l<strong>and</strong>scape permeability, we applied graph<br />

theory in order to highlight the main habitat patches influencing habitat availability <strong>and</strong><br />

connectivity by the use of the software’s Conefor Sensinode 2.2 <strong>and</strong> Guidos. The use of the JRC<br />

<strong>Forest</strong>/Non <strong>Forest</strong> European map for the characterisation of common frog terrestrial habitat<br />

distribution combined with the maximum entropy modelling gives promising results for the<br />

identification of habitat discontinuities within a regional perspective. This approach should<br />

provide an operational tool for the identification of the effects of “l<strong>and</strong>scape barriers <strong>and</strong><br />

corridors” on populations structure. Then, the method appears as a promising tool for l<strong>and</strong>scape<br />

planning.<br />

Key words: common frog, l<strong>and</strong>scape connectivity, habitat suitability modelling, graph theory,<br />

maximum entropy modelling<br />

1. Introduction<br />

L<strong>and</strong>scape connectivity is considered a key issue for biodiversity conservation <strong>and</strong> for the<br />

maintenance of natural ecosystems stability <strong>and</strong> integrity. L<strong>and</strong>scape connectivity defines the<br />

degree to which the l<strong>and</strong>scape facilitates or impedes movement among resource patches (Taylor<br />

<strong>and</strong> al. 1993). In fragmented <strong>and</strong> heterogeneous human dominated l<strong>and</strong>scapes, movements<br />

across the l<strong>and</strong>scape matrix area are key process for the survival of plant <strong>and</strong> animals species.<br />

Maintaining or restoring l<strong>and</strong>scape connectivity has become a major concern in conservation<br />

biology <strong>and</strong> l<strong>and</strong> planning (Pascual-Hortal <strong>and</strong> Saura 2008) <strong>and</strong> especially for amphibians.<br />

Indeed, amphibian’s life cycle involves seasonal migrations between terrestrial <strong>and</strong> aquatic<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Decout et al. 2010. Connectivity loss in human dominated l<strong>and</strong>scape<br />

662<br />

habitats which constrain them to regularly cross an inhospitable fragmented l<strong>and</strong>scape matrix<br />

making them vulnerable to l<strong>and</strong> degradation <strong>and</strong> connectivity loss (Allentoft <strong>and</strong> O’Brien 2010).<br />

Anthropogenic barriers as railways <strong>and</strong> major roads limit amphibians’migrations <strong>and</strong><br />

movements. Many species have to refrain to move between small, scattered patches of different<br />

resources, instead of one, large patch. In this sense, habitat fragmentation constitutes the main<br />

driver of gene flow reduction (Allentoft <strong>and</strong> O’Brien 2010). This is particularly the case for the<br />

common frog Rana temporaria, a widespread amphibian in Europe occurring in various habitat<br />

types <strong>and</strong> migrating between forest <strong>and</strong> aquatic habitats for breeding (Miaud <strong>and</strong> al. 1999). The<br />

study focus on habitat availability <strong>and</strong> l<strong>and</strong>scape connectivity, under the assumption that<br />

connectivity is species specific <strong>and</strong> should be measured from a functional perspective (Saura<br />

<strong>and</strong> Torné 2009). The focus is on viable habitat patches, in relation to the ongoing need for a<br />

holistic approach to l<strong>and</strong>scapes <strong>and</strong> habitats. The overall goal is to find a continuum for habitat<br />

suitability of the species in question. Graph theory <strong>and</strong> network analysis have become<br />

established as promising ways to efficiently explore <strong>and</strong> analyze l<strong>and</strong>scape or habitat<br />

connectivity. However, little attention has been paid to making these graph-theoretic approaches<br />

operational within l<strong>and</strong>scape ecological assessments, planning, <strong>and</strong> design. We are working<br />

towards a methodological approach to address habitat quality assessment <strong>and</strong> connectivity from<br />

an operational point of view in order to support planning. To illustrate the basic principles of the<br />

proposed method, an ecological example using the European common frog Rana temporaria, in<br />

the French Alps region is presented. The approach is based on three main steps: i) Achievement<br />

of a probability of occurrence distribution map by the use of presence data <strong>and</strong> maximum<br />

entropy modelling ii) Simulation of dispersal areas in order to define the main connections<br />

between common frog ponds iii) Assessment of the main connected ponds by the use of graph<br />

theory <strong>and</strong> the software Conefor Sensinode 2.2 (Saura <strong>and</strong> Torné, 2009). We present here<br />

preliminary results on undergoing research, in order to exchange ideas in relation with this<br />

approach.<br />

2. Methodology<br />

2.1 Study site <strong>and</strong> sampling<br />

This study focuses on the French departments Isère <strong>and</strong> Savoie (French Alps). This area is about<br />

1415126 km² (see figure 1). The common frog is a typical species within this region where it<br />

breeds in various types of aquatic habitats. Because at the subalpine belt l<strong>and</strong>scape connectivity<br />

is not the main driver of the frog dispersal patterns due to environmental constraints (i.e.<br />

climatic variables), we focused on the common frog populations occurring under the tree line<br />

(1400-1600 meters). The common frog was detected in 97 ponds under the tree line within this<br />

area. The sample design followed a genetic sampling strategy framework based on tadpoles<br />

between 1999 <strong>and</strong> 2002 (Pidancier <strong>and</strong> al. 2002). The geographic location of each sampling is<br />

known. For this preliminary study, we reduced the area to a surface of 4067 km² including 47<br />

located ponds (see figure 1).<br />

Figure 1: The study area.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Decout et al. 2010. Connectivity loss in human dominated l<strong>and</strong>scape<br />

663<br />

2.2 Probability of occurrence distribution<br />

We considered the 47 genetic sampling locations as presence data. It must be noted that the<br />

approach is based in present information only of the common frog within the study area. We<br />

used the maximum entropy modelling approach. In order to assess the distribution of the<br />

probability of detection of the common frog, we used in particular the software MaxEnt (Philips<br />

<strong>and</strong> Dudik 2008). Different environmental variables were analyzed <strong>and</strong> used to develop the<br />

probability of occurrence distribution map with MaxEnt. The common frog during its terrestrial<br />

cycle is very sensitive to the type of l<strong>and</strong> cover to cross in order to reach its required forest<br />

habitat for summer <strong>and</strong> winter (Miaud <strong>and</strong> al. 1999). Based on radiotracking surveys <strong>and</strong> expert<br />

knowledge, the common frog seems to be very sensitive to the distribution of small forest<br />

patches around the pond area. Consequently, we computed <strong>and</strong> integrated in the analysis<br />

different environmental variables in relation to ecological <strong>and</strong> spatial requirements of the<br />

common frog. The forest habitat distribution around the aquatic habitat was also considered<br />

within the modelling:<br />

1. L<strong>and</strong> cover based on Corine L<strong>and</strong> Cover 2006 (level 3).<br />

2. Slope <strong>and</strong> elevation derived from a 50m DEM (French National Geographic Institute).<br />

3. L<strong>and</strong>scape indices based on forest patches distribution from the European <strong>Forest</strong>/Non<br />

<strong>Forest</strong> map (resolution: 25m) provided by the Join Research Centre JRC. For this, we<br />

used Fragstats (McGarigal <strong>and</strong> Marks 1995) with a moving window of 3000m <strong>and</strong> we<br />

selected the following basics l<strong>and</strong>scape indices: Mean <strong>Forest</strong> Patch Area, Largest <strong>Forest</strong><br />

Patch Index <strong>and</strong> <strong>Forest</strong> Patch Density.<br />

4. Distance to forest patches crossed by a river derived from a combination of the<br />

hydrological network map (French National Geographic Institute) with the European<br />

<strong>Forest</strong>/Non <strong>Forest</strong> map.<br />

2.3 Connections between ponds<br />

We quantified the connection between the ponds in relation with l<strong>and</strong>scape matrix permeability<br />

by the use of a friction map <strong>and</strong> the least cost modelling. Least cost modelling allows to<br />

simulate the dispersal of the common frog in relation to the l<strong>and</strong>scape matrix permeability<br />

between habitat patches. The matrix permeability is considered with the use of a friction map<br />

that provides inputs in terms of the ability of the common frog to cross the l<strong>and</strong>scape matrix.<br />

The friction map layer is a raster map where each cell (l<strong>and</strong>scape unit) expresses the relative<br />

difficulty of moving through that cell (Fulgione <strong>and</strong> al. 2009). In this study, the present friction<br />

map was computed by inversing the previous probability of occurrence distribution map from<br />

MaxEnt (Fulgione <strong>and</strong> al. 2009). Indeed, a fundamental assumption is that habitat suitability<br />

<strong>and</strong> permeability are synonyms, <strong>and</strong> that both are the inverse of ecological cost of travel (Beier<br />

<strong>and</strong> al. 2007). We added the highways <strong>and</strong> the urbanized areas to this friction map in order to<br />

integrate the main “impermeable barriers” for the common frog (i.e. high friction value). For the<br />

calculation of the least cost paths between each pond, we used the ArcView extension Path<br />

Matrix (Ray 2005).<br />

2.4 Assessment of ponds’ importance for connectivity<br />

We considered all the located ponds as nodes in order to use graph theory, in particular Conefor<br />

Sensinode software (Saura <strong>and</strong> Torné 2009). The least cost paths distances between the ponds<br />

allowed to calculate a set of quantitative connectivity rules between ponds. The software<br />

calculates a Number of Components NC index which identify a set of connected nodes (i.e.<br />

components) in which a path exists between every pair of nodes. The software also allows to<br />

calculate a Probability of Connectivity index (PC), which combines the attribute of the nodes<br />

with the maximum product probability of all the possible paths between every pair of nodes<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Decout et al. 2010. Connectivity loss in human dominated l<strong>and</strong>scape<br />

664<br />

(Saura <strong>and</strong> Torné 2009). All the more, the software provides to assess node importance for<br />

connectivity by removing systemically each node <strong>and</strong> recalculating the PC when that node is<br />

not present in the l<strong>and</strong>scape. Node importance is quantified by an index dPC which corresponds<br />

to the importance of an existing node for maintaining l<strong>and</strong>scape connectivity according to the<br />

PC index variation when the node is removed (Pascual-Hortal <strong>and</strong> Saura 2008). In our case<br />

study, we used a threshold dispersal distance of 1500m based on radiotracking surveys of<br />

common frog migration pattern between ponds <strong>and</strong> suitable terrestrials’ habitats.<br />

3. Results<br />

3.1 Probability of occurrence distribution map <strong>and</strong> habitat suitability map<br />

The use of 15% of the dataset for cross validation gives an Area under the Curve (AUC) of 0.75<br />

for the ROC curve analysis which corresponds to a good discriminative capability between<br />

predicted presence <strong>and</strong> absence according to Pearce <strong>and</strong> Ferrier (2000). The figure 2 shows the<br />

resulting common frog probability of occurrence distribution map. The environmental variables<br />

with highest gain when used in isolation are Elevation <strong>and</strong> Largest <strong>Forest</strong> Patch Index in the<br />

jackknife test of variable importance in MaxEnt. MaxEnt also calculates several threshold<br />

values at each run <strong>and</strong> values exceeding them may be interpreted as reasonable approximation<br />

of the potential distribution of the considered species suitable habitat. As suggested by Phillips<br />

<strong>and</strong> Dudik (2008), we used the 10 percentile training presence (mean = 0.339) in order to obtain<br />

the potential distribution of the common frog in relation to suitable terrestrial habitat<br />

distribution (see figure 2). The potential distribution of the common frog obtained (see figure 2)<br />

allows to identify the effect of the dense urbanized areas <strong>and</strong> highways as main barriers <strong>and</strong><br />

unsuitable habitats. This distribution also suggests the potential presence of discontinued<br />

potential suitable areas for the frog depending on forest patches distribution impacted by human<br />

activities. In this context, further genetic considerations will help to quantify <strong>and</strong> identify the<br />

disconnections between frog populations in relation to human dominated areas distribution.<br />

Figure 2: Probability of occurrence distribution for the common frog with the maximum entropy<br />

modelling <strong>and</strong> resulting potential distribution of the common frog (10 percentile training presence of<br />

0.339 as the probability threshold) (area of 4067 km²).<br />

3.2 Ponds’ importance for connectivity<br />

The use of the NC index (see figure 3) provides a rapid identification of the connected ponds in<br />

relation with l<strong>and</strong>scape matrix. In our case study, most of the ponds are isolated by distance <strong>and</strong><br />

few ponds can be considered as connected in term of seasonal migration patterns. All the more,<br />

most of the connected ponds identified are located in homogenous suitable habitat. This is due<br />

to the orientation of the ponds location dataset for genetic analysis (genetic isolation by<br />

distance). Within this context, we plan to improve the analysis using a more detailed pond’<br />

distribution dataset in order to assess local connectivity in the near future. On figure 3, some<br />

ponds isolated <strong>and</strong> closed to urbanized areas appear as important for regional connectivity (high<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Decout et al. 2010. Connectivity loss in human dominated l<strong>and</strong>scape<br />

665<br />

dPC value). This may suggest that these ponds could be considered as critical isolated ponds in<br />

relation with barriers in a human dominated l<strong>and</strong>scape context (presence of disconnections<br />

between suitable large areas for the common frog). For the moment, this interpretation of the<br />

dPC has to be considered with caution given that we did not use yet all the exisiting ponds<br />

locations within the area (missing nodes). We plan to complete the study with the computation<br />

of a dPC index based on genetic distance between ponds for the quantification of the potential<br />

genetic connections.<br />

Figure 3: Set of connected ponds (components) identified with the computation of the Number of<br />

Components index (NC) <strong>and</strong> ponds importance for connectivity based on the computation of the dPC<br />

index (warmer colours correspond to a highest importance for connectivity).<br />

4. Discussion<br />

In this preliminary study, the use of the JRC <strong>Forest</strong>/Non <strong>Forest</strong> European map for the<br />

characterisation of common frog terrestrial habitat distribution combined with the maximum<br />

entropy modelling gives promising results for the identification of discontinuities in distribution<br />

within a regional perspective. This approach in t<strong>and</strong>em with genetic considerations should<br />

provide a tool for the identification of the effects of “l<strong>and</strong>scape barriers <strong>and</strong> corridors” on<br />

populations structure in relation to common frog <strong>and</strong> its terrestrial habitat requirements. The use<br />

of a friction map combined with least path modelling appears also as a crucial key issue for the<br />

quantification of connections between habitat patches when dealing with l<strong>and</strong>scape matrix<br />

permeability. Even if an efficient calibration of a friction map is possible for a local approach<br />

(Janin <strong>and</strong> al. 2009), the computation of a relevant regional friction map remains quite difficult<br />

for the common frog given the existence of heterogeneity in dispersal patterns driven by local<br />

environmental conditions. This suggests that it should be more efficient to consider regional<br />

connectivity for amphibians from the point of view of genetic <strong>and</strong> spreading diseases as the<br />

chytrid fungus (Rödder <strong>and</strong> al 2009). L<strong>and</strong>scape connectivity should be better considered for a<br />

local perspective in relation with common frog migration patterns between its aquatic <strong>and</strong><br />

terrestrial habitats. In this context, the use of a graph theoretical approach appears as a<br />

promising tool for the assessment of local l<strong>and</strong>scape connectivity for the common frog given<br />

that the Conefor Sensinode software provides a powerful tool integrating considerations about<br />

habitat patches distribution <strong>and</strong> suitability in a l<strong>and</strong>scape matrix context surrounding habitat<br />

patches. Moreover is proven as an operational tool to identify barriers <strong>and</strong> important patches for<br />

planning purposes.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Decout et al. 2010. Connectivity loss in human dominated l<strong>and</strong>scape<br />

666<br />

Acknowledgements<br />

This research was supported <strong>and</strong> funded by the Interreg Alpine Space Program Econnect<br />

(reference number: 116/1/3/A). We thank Santiago Saura for his advice on the subject <strong>and</strong> his<br />

help on using the software Conefor Sensinode.<br />

References<br />

Allentoft, M.E. <strong>and</strong> O’Brien, J., 2010. <strong>Global</strong> amphibian declines, loss of genetic diversity <strong>and</strong><br />

fitness: a review. Diversity, 47-71.<br />

Beier, P., Majka, D., Jenness, J., 2007. Conceptual steps for designing wildlife corridor.<br />

www.corridordesign.org, 86 p.<br />

Fulgione, D., Maselli, V., Pavarese, G., Rippa, D., Rastogi, R.K., 2009. L<strong>and</strong>scape<br />

fragmentation <strong>and</strong> habitat suitability in endangered Italian hare (Lepus corsicanus) <strong>and</strong><br />

European hare (lepus europaeus) populations. European Journal on Wildlife Research,<br />

55:3875-396.<br />

Janin, A., Léna, J.-P., Ray, N., Delacourt, C., Allem<strong>and</strong>, P., Joly, P., 2009. Assessing l<strong>and</strong>scape<br />

connectivity with calibrated cost-distance modelling: predicting common toad distribution in a<br />

context of spreading agriculture. Journal of Applied Ecology, 1-9.<br />

McGarigal, K. <strong>and</strong> Mark,s B.J. 1995. FRAGSTATS: Spatial Pattern Analysis Program for<br />

Quantifying L<strong>and</strong>scape Structure. General Technical Report PNW-GTR-351. Pacific Northwest<br />

research Station, <strong>Forest</strong> Service, US Department of Agriculture, Portl<strong>and</strong>, OR.<br />

Miaud, C., Guyétant, R., Elmberg, J., 1999. Variation in life-history traits in the common frog<br />

Rana temporaria (Amphibia: Anura): a literature reviews <strong>and</strong> new data from the French Alps.<br />

Journal of Zoology, 249: 61-73.<br />

Pascual-Hortal, L. <strong>and</strong> Saura S., 2008. Integrating l<strong>and</strong>scape connectivity in broad-scale forest<br />

planning through a new graph-based habitat availability methodology: application to<br />

capercaillie (Tetrao urogallus) in Catalonia (NE Spain). European Journal of <strong>Forest</strong> Research,<br />

127: 23-31.<br />

Pearce, J. <strong>and</strong> Ferrier, S., 2000. Evaluating the predictive performance of habitat models<br />

deleveloped using logistic regression. Ecological Modelling, 133:225-245.<br />

Pidancier, N., Gauthier, P., Miquel, C., Pompanon, F., 2002. Polymorphic microsatellite DNA<br />

loci identified in the common frog (Rana temporaria, Amphibia, Ranidae). Molecular Ecology<br />

Notes, 3: 304-305.<br />

Philips, S.J. <strong>and</strong> Dudik, M., 2008. Modeling of species distributions with Maxent: new<br />

extensions <strong>and</strong> a comprehensive evaluation. Ecography, 31: 161-175.<br />

Ray, N. 2005. PATHMATRIX: a geographical information system tool to compute effective<br />

distances among samples. Molecular Ecology Notes, 5: 177–180.<br />

Rödder, D., Kielgast, J., Bielby, J., Schmidtlein, S., Bosch, J., Garner, T., Veith, M., Walker, S.,<br />

Fisher, M., Lötters, S., 2009. <strong>Global</strong> amphibian risk assessment for the panzootic chytrid<br />

fungus. Diversity, 1: 52-66.<br />

Riitters, K.H., Wickham, J.D., O’Neill, R.V., Jones, K.B., Smith, E.R., 2000. <strong>Global</strong>-scale<br />

patterns of forest fragmentation. Conservation Ecology, 4(2): 3.<br />

Riitters, K., Vogt, P., Soille, P., Estreguil, C., 2009. L<strong>and</strong>scape petterns from mathematical<br />

morphology on maps with contagion. L<strong>and</strong>scape Ecology, 24: 699-709.<br />

Saura, S., Torné, J., 2009. Conefor Sensinode 2.2: A sofware package for quantifying the<br />

importance of habitat patches for l<strong>and</strong>scape connectivity. Environmental Modelling <strong>and</strong><br />

Sofware, 24:135-139.<br />

Taylor, P.D., Fahrig, L., Henein, K., Merriam, N.G., 1993. Connectivity is a vital element of<br />

l<strong>and</strong>scape structure. Oikos, 68: 571-573.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


L<strong>and</strong>scape genetics


S. Joost et al. 2010. GEOME. Towards an integrated web-based l<strong>and</strong>scape genomics platform<br />

668<br />

GEOME<br />

Towards an integrated web-based l<strong>and</strong>scape genomics platform<br />

Stéphane Joost 1* , Michael Kalbermatten 1 & Nicolas Ray 2<br />

1<br />

GIS Research Laboratory (LASIG), Institute of Environmental Engineering (IIE),<br />

School of Architecture, Civil <strong>and</strong> Environmental Engineering (ENAC), Ecole<br />

Polytechnique Fédérale de Lausanne (EPFL), Switzerl<strong>and</strong><br />

2<br />

EnviroSPACE Laboratory, Institute for environmental Sciences (ISE),<br />

University of Geneva, Switzerl<strong>and</strong><br />

Abstract<br />

The aim of the GEOME project is to develop a WebGIS-based platform for the integrated<br />

analysis of environmental, ecological <strong>and</strong> molecular data through the implementation of an<br />

original set of combined databases, spatial analysis <strong>and</strong> population genetics tools, in a High<br />

Performance Computing context. GEOME will support tasks of l<strong>and</strong>scape ecologists <strong>and</strong><br />

resource conservation managers who increasingly have to use geo-referenced data but are not<br />

trained to efficiently use Geographical Information Systems together with appropriate geoenvironmental<br />

information <strong>and</strong> spatial analysis approaches. Preliminary developments (three<br />

first applications) are already available here: http://lasigpc8.epfl.ch/geome/<br />

Keywords: l<strong>and</strong>scape genomics, geocomputation, environmental data, adaptation, GIS<br />

1. Introduction<br />

Since 2000, we are witnessing a progressive integration of the fields of ecology, evolution <strong>and</strong><br />

population genetics (Lawry 2010). The combination of l<strong>and</strong>scape ecology with population<br />

genetics led to the advent of l<strong>and</strong>scape genetics whose goal is to underst<strong>and</strong> how geographical<br />

<strong>and</strong> environmental features structure genetic variation in living organisms (Manel 2003).<br />

Recently, l<strong>and</strong>scape genomics (Luikart 2003; Joost 2007; Lawry 2010) emerged as a research<br />

field at the interface of genome sciences, environmental resources analysis <strong>and</strong> spatial statistics.<br />

The combination of these fields permits to assess the level of association between specific<br />

genomic regions <strong>and</strong> environmental factors to identify loci responsible for adaptation to<br />

different habitats (Lawry 2010). Henceforth, knowledge of geo-environmental data <strong>and</strong> skills in<br />

GIS are a necessity to develop research or management activities in these l<strong>and</strong>scape sciences.<br />

To favor the use of l<strong>and</strong>scape genomics <strong>and</strong> to facilitate access to necessary geo-environmental<br />

data, GIS <strong>and</strong> spatial analysis tools, the development of a robust <strong>and</strong> efficient computer<br />

infrastructure is required. Therefore, GEOME will constitute a robust, easy-to-use <strong>and</strong> powerful<br />

internet platform based on High Performance Computing (HPC), able to h<strong>and</strong>le <strong>and</strong> process<br />

very large genome <strong>and</strong> environmental data sets, <strong>and</strong> offering facilities for the statistical <strong>and</strong><br />

(geo)visual analysis of results (http://lasigpc8.epfl.ch/geome/).<br />

The GEOME platform will support access to free geo-environmental data (database) <strong>and</strong><br />

provide dedicated GIS tools to scientists <strong>and</strong> professional users (e.g. l<strong>and</strong>scape ecologists or<br />

resource conservation managers) who often do not have a background nor a training in<br />

* Corresponding author. Tel.:+41216935782 - Fax:+41216935790<br />

Email address: stephane.joost@epfl.ch<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Joost et al. 2010. GEOME. Towards an integrated web-based l<strong>and</strong>scape genomics platform<br />

669<br />

GIScience (Geographic Information Systems <strong>and</strong> spatial analysis), <strong>and</strong> do not have time to<br />

acquire them. To be able to carry out their investigations, these users need support to i) find<br />

appropriate geo-referenced environmental data sets to address their research problematics, ii)<br />

integrate their own data sets (e.g. presence/absence of genotypes, of species) with the geoenvironmental<br />

information mentioned here above, iii) benefit from specific analytical tools<br />

implemented within a simple GIS environment.<br />

Several web-based platforms already exist in the domain of genomics or other genetic topics<br />

(population genetics, phylogenetics). An interesting example is the BIOPORTAL platform<br />

(http://www.bioportal.uio.no), a web-based service platform developed at University of Oslo for<br />

phylogenomic analysis, population genetics <strong>and</strong> high-throughput sequence analysis.<br />

BIOPORTAL is the largest publicly available computer resource with 300 dedicated<br />

computational cores in a cluster named TITAN. Currently, applications in chemistry <strong>and</strong><br />

economics have been integrated, but other applications will be added along the way. Another<br />

example is GeneGrid, a web-based collaborative industrial grid computing solution developed<br />

in the United Kingdom. It was initiated by the Belfast e-Science Centre (BeSC) under the UK e-<br />

Science programme <strong>and</strong> supported by the UK Department of Trade & Industry (DTI). This<br />

platform gives access to collective resources, skills, experiences <strong>and</strong> results in a secure, reliable<br />

<strong>and</strong> scalable manner through the creation of a “Virtual Bioinformatics Laboratory”. GeneGrid<br />

provides seamless integration of a series of heterogeneous applications <strong>and</strong> data sets that span<br />

multiple administrative domains <strong>and</strong> locations across the globe, <strong>and</strong> presents these to the<br />

scientist through a simple user friendly interface (Kelly et al. 2005).<br />

It has to be noted that no existing Web-based platform includes the combination of services to<br />

be implemented within GEOME. This combination will permit to address the issue of local<br />

adaptation through the complementary implementation of a theoretical approach in population<br />

genomics (Foll <strong>and</strong> Gaggiotti 2008) <strong>and</strong> of a spatial analysis approach (Joost et al. 2007).<br />

2. Main issue addressed: local adaptation<br />

In the present context of rapid global climate change, ecologists <strong>and</strong> evolutionists show a<br />

renewed interest to study adaptation (Holderegger et al. 2008). Local adaptation in particular is<br />

an important issue in conservation genetics. For example, this process has to be better<br />

understood in order to correctly consider the effects of transfers of individuals between<br />

populations, which is often a technique proposed to replenish genetic variation <strong>and</strong> to reduce<br />

negative effects of low genetic diversity. The study of local adaptation is also likely to provide<br />

objective <strong>and</strong> unambiguous criteria to characterize conservation areas – in combination with<br />

models of predictive habitat or Species Distribution Models (SDMs, Guisan <strong>and</strong> Thuiller 2005)<br />

– which are the most worthwhile preserving.<br />

In parallel, development of low impact sustainable agriculture as well as husb<strong>and</strong>ry based on<br />

adapted breeds is of priority to most countries in the world, <strong>and</strong> is of key importance to<br />

emerging countries in particular. The genetic basis <strong>and</strong> the level of adaptation of livestock<br />

breeds to their environment has to be investigated, in order to reach better underst<strong>and</strong>ing of the<br />

relationship between environment <strong>and</strong> the adaptive fitness of livestock populations, <strong>and</strong> to favor<br />

sustainable production systems based on adapted breeds.<br />

During the last decade, “tremendous advances in genetic <strong>and</strong> genomic techniques have resulted<br />

in the capacity to identify genes involved in adaptive evolution across numerous biological<br />

systems” (Lowry 2010). There is now an important need to provide tools allowing the<br />

acceleration of the study of how l<strong>and</strong>scape-level geographical <strong>and</strong> environmental features are<br />

involved in the distribution of functional adaptive genetic variation, as highlighted by recent<br />

publications (Lowry 2010). A few years ago, Holderegger <strong>and</strong> Wagner (2008) already stated<br />

that “Novel approaches linking spatially explicit environmental analysis with molecular<br />

genetics could offer effective means to study the spread of adaptive genes across l<strong>and</strong>scapes”.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Joost et al. 2010. GEOME. Towards an integrated web-based l<strong>and</strong>scape genomics platform<br />

670<br />

L<strong>and</strong>scape genomics is one of these approaches, <strong>and</strong> GEOME will facilitate its use by means of<br />

a database providing access to the many geo-environmental data sets available worldwide.<br />

3. Access to remote environmental data<br />

Joost et al. 2010 provided a non-exhaustive list of the many environmental data sets available<br />

on the Internet. Different initiatives at the regional <strong>and</strong> global levels influence <strong>and</strong> promote the<br />

creation of Spatial Data Infrastructures (SDIs) to access these data. They also work on the<br />

harmonization, st<strong>and</strong>ardization, interoperability, <strong>and</strong> seamless integration of the different GIS<br />

layers constituting these data. An example is the <strong>Global</strong> Earth Observation System of Systems<br />

(GEOSS), which is a worldwide effort to connect already existing SDIs <strong>and</strong> Earth Observation<br />

infrastructures. Through its developing GEO portal <strong>and</strong> related Common Infrastructure, GEOSS<br />

is foreseen to act as a gateway between producers of environmental data <strong>and</strong> end users, with the<br />

aim of enhancing the relevance of Earth observations for global issues <strong>and</strong> to offer public access<br />

to comprehensive information <strong>and</strong> analyses on the environment (GEO secretariat 2007).<br />

Today's effort on the technical development of SDI components clearly focuses on the exchange<br />

of geodata in an interoperable way (Bernard <strong>and</strong> Craglia 2005), which is highlighted by the<br />

concept of web services <strong>and</strong> the related Service Oriented Architecture (SOA). Web services<br />

constitute a “new paradigm" allowing users to retrieve, manipulate <strong>and</strong> combine geospatial data<br />

from different sources <strong>and</strong> different formats using HTTP protocol to communicate (Sahin <strong>and</strong><br />

Gumusay 2008). Web services enable the possibility to construct web-based application using<br />

any platform, object model <strong>and</strong> programming language. The Open Geospatial Consortium<br />

(OGC) has specified a suite of st<strong>and</strong>ardized web services. Two of them are of particular interest<br />

for data providers <strong>and</strong> users: the Web Feature Service (WFS) that provides a web interface to<br />

access vector geospatial data (like country borders, GPS points or roads) encoded in Geographic<br />

Markup Language (GML) <strong>and</strong> the Web Coverage Service (WCS) that defines a web interface to<br />

retrieve raster geospatial data of spatially distributed phenomena such as surface temperature<br />

maps or digital elevation models (DEMs). In addition, the Web Map Service (WMS) defines an<br />

interface to serve georeferenced map images suitable for displaying purpose based on either<br />

vector or raster data. GEOME’s Web services will provide the geocomputational context with<br />

the support of an indispensable High Performance Computing (HPC) infrastructure to enable<br />

the processing of associations models between millions of loci (see section 4) <strong>and</strong> hundreds of<br />

environmental parameters (see section 5).<br />

4. GEOME’s challenge: whole genome scan<br />

One of the next major steps in evolutionary biology is to determine how l<strong>and</strong>scape-level<br />

geographical <strong>and</strong> environmental features are involved in the distribution of the functional<br />

adaptive genetic variation (Lawry 2010). This challenge will take place in a context where the<br />

amount of molecular data to be analyzed will exp<strong>and</strong> very rapidly. Indeed, after the long (13<br />

years) <strong>and</strong> expensive (3 billion dollars) human genome sequencing project (completed in 2003),<br />

the American National Institutes of Health (NIH) proposed in 2004 the challenge to sequence<br />

one human genome for $1’000 (Service 2006). And future generation of sequencers will allow<br />

researchers to get to genomic data faster <strong>and</strong> at a lower cost. For example, the theoretical<br />

potential of single-molecule/nanopore sequencing is undeniable (Tersoff 2001). Based on this<br />

technology, with a 100 nanopores in parallel, a mammalian genome could be sequenced in 24<br />

hours with the main cost being the chip itself, probably around $1'000 (Blow 2008). Several<br />

alternative low cost sequencing technologies are under way, even decreasing to $100 in the case<br />

of the Pacific Biosciences technology (Eid et al. 2009).<br />

Thus, any research project in l<strong>and</strong>scape genomics will very soon be given the opportunity to<br />

investigate the entire genome of sampled individuals, meaning that from now on HPC is<br />

required to process these huge quantities of data when compared to eco-climatic parameters. To<br />

be ready to h<strong>and</strong>le such volumes of data, the GIS laboratory (LASIG) involved in the<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Joost et al. 2010. GEOME. Towards an integrated web-based l<strong>and</strong>scape genomics platform<br />

671<br />

development of GEOME is also a partner within the EU FP7 NEXTGEN, the first project in the<br />

area of conservation genetics that proposes a comparative analysis of whole genome data at the<br />

intraspecific level (http://nextgen.epfl.ch). NEXTGEN will also need GEOME’s expertise in the<br />

area of remote sensing, digital elevation models <strong>and</strong> other environmental data in general.<br />

5. Geo-environmental data in a HPC environment<br />

Environmental sciences are a data-intensive domain in which applications typically produce <strong>and</strong><br />

analyze a large amount of geospatial data. Moreover, due to the multi-disciplinary nature of<br />

environmental sciences (e.g. ecology, climate change, etc.), scientists need to integrate large<br />

amount of data distributed all around the world in different data centers. A local cluster is the<br />

first mean to upscale computational capacities when the workflow can be distributed into many<br />

independent jobs.<br />

When the number of jobs is very large <strong>and</strong>/or users do not belong to the Institution owning the<br />

cluster, grid computing can be an efficient solution. Following Foster et al. (2008), a grid is a<br />

parallel processing architecture in which computational resources are shared across a network<br />

allowing access to unused CPU <strong>and</strong> storage space to all participating machines. Resources could<br />

be allocated on dem<strong>and</strong> to consumers who wish to obtain computing power. Recent studies have<br />

had a successful approach to extend grid technology to the remote sensing community (Muresan<br />

et al. 2006), as well as in the field of disaster management (Mazzetti et al. 2009) making OGC<br />

web service grid-enabled.<br />

One of the largest scientific grid infrastructures currently in operation is the Enabling Grids for<br />

E-sciencE (EGEE) infrastructure bringing together more than 120 organisations to provide<br />

scientific computing resources to the European <strong>and</strong> global research community. The currently<br />

120'000 CPUs available in EGEE are essentially used by the High Energy physics community,<br />

but part of EGEE is also opened to other disciplines such as environmental sciences or genetics.<br />

A grid environment federates its users through Virtual Organizations (VOs), which are sets of<br />

individuals <strong>and</strong>/or institutions defined by a set of sharing rules (e.g. access to computers,<br />

software, data <strong>and</strong> other resources). One particular VO of interest is named Biomed; it has<br />

currently access to 20'000 CPUs <strong>and</strong> is willing to accept the envisioned GEOME on the Biomed<br />

VO.<br />

6. Conclusion<br />

No software tool presently exists to answer challenges of developing better approaches for<br />

linking ecologically relevant data sets to specific loci or genes. Moreover, no software tool can<br />

currently integrate geo-environmental variables <strong>and</strong> molecular data within a WebGIS<br />

environment, with analysis modules included i) to detect loci under selection according to two<br />

complementary approaches, ii) to analyze <strong>and</strong> characterize the spatial distribution of these loci,<br />

<strong>and</strong> iii) to produce predictive habitat maps derived from them.<br />

Thanks to a robust HPC infrastructure, GEOME’s existing <strong>and</strong> future Web services will be<br />

useful to a very large number of users, <strong>and</strong> will possibly contribute in underst<strong>and</strong>ing how<br />

l<strong>and</strong>scape-level geographical <strong>and</strong> environmental features are involved in the distribution of the<br />

functional adaptive genetic variation.<br />

References<br />

Bernard, L., <strong>and</strong> Craglia, M., 2005. SDI - From Spatial Data Infrastructure to Service Driven<br />

Infrastructure. First Research Workshop on Cross-learning on Spatial Data<br />

Infrastructures (SDI) <strong>and</strong> Information Incrastructures (II), Enschende, The Netherl<strong>and</strong>s.<br />

Blow, N., 2008. DNA sequencing: generation next-next. Nature Methods, 3: 267- +.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. Joost et al. 2010. GEOME. Towards an integrated web-based l<strong>and</strong>scape genomics platform<br />

672<br />

Eid, J. et al., 2009. Real-Time DNA Sequencing from Single Polymerase Molecules. Science,<br />

5910: 133-138.<br />

Foster, I, Zhao, Y., Raicu, I., Lu, S., 2008. Cloud Computing <strong>and</strong> Grid Computing 360-degree<br />

compared. IEEE Grid Computing Environments (GCE08), pp. 1–10.<br />

Guisan, A. <strong>and</strong> Thuiller, W., 2005. Predicting species distribution: offering more than simple<br />

habitat models. Ecology Letters, 8: 993-1009.<br />

GEO secretariat, 2007. Strategic Guidance for Current <strong>and</strong> Potential Contributors to GEOSS.<br />

Fourth Plenary Session of GEO (GEO-IV), Cape Town, 4 p.<br />

http://www.earthobservations.org/docs_od_ple.shtml<br />

Holderegger, R. <strong>and</strong> Wagner, H.H. (2008) L<strong>and</strong>scape genetics. Bioscience, 58: 199-207.<br />

Holderegger, R., Herrmann, D., Poncet, D. et al., 2008. L<strong>and</strong> ahead: using genome scans to<br />

identify molecular markers of adaptive relevance. Plant Ecology & Diversity, 1: 273-283.<br />

Joost, S., Bonin, A., Bruford, M.W., Després, L., Conord, C., Erhardt, G., Taberlet, P., 2007. A<br />

Spatial Analysis Method (SAM) to detect c<strong>and</strong>idate loci for selection: towards a<br />

l<strong>and</strong>scape genomics approach to adaptation. Molecular Ecology, 18: 3955–3969.<br />

Joost, S., Colli, L., Baret, P.V., Garcia, J.F., Boettcher, P.J., Tixier-Boichard, M. , Ajmone-<br />

Marsan, P. & the GLOBALDIV Consortium, 2010. Integrating geo-referenced multiscale<br />

<strong>and</strong> multidisciplinary data for the management of biodiversity in livestock genetic<br />

resources. Animal Genetics, 41:47-63.<br />

Kelly, N., Jithesh, P.V., Donachy, P., Harmer, T.J., Perrott, R.H. (2005) GeneGrid: A<br />

Commercial Grid Service Oriented Virtual Bioinformatics Laboratory. Proceedings of the<br />

2005 IEEE International Conference on Services Computing (SCC’05), July 11-15,<br />

Orl<strong>and</strong>o, 1:42-50.<br />

Manel, S., Schwartz, M., Luikart, G., Taberlet, P., 2003. L<strong>and</strong>scape genetics: combining<br />

l<strong>and</strong>scape ecology <strong>and</strong> population genetics. Trends in Ecology & Evolution, 18: 189-197.<br />

Mazzetti, P., S. Nativi, Angelini, V. Verlato, M., Fiorucci, P., 2009. A Grid platform for the<br />

European Civil Protection e-Infrastructure: the <strong>Forest</strong> Fires use scenario. Earth Science<br />

Informatics, 2: 53-62.<br />

Muresan, O., Pop, F., Gorgan, D., Cristea, V., 2006. Satellite Image Processing Applications in<br />

MedioGRID. Proceedings of The Fifth International Symposium on Parallel <strong>and</strong><br />

Distributed Computing (ISPDC'06), 6-9 July 2006, Timisoara, pp.253-262.<br />

Lowry, D.B., 2010. L<strong>and</strong>scape evolutionary genomics. Biology Letters, DOI:<br />

10.1098/rsbl.2009.0969.<br />

Luikart, G., Engl<strong>and</strong>, P.R., Tallmon, D., Jordan, S., Taberlet, P., 2003. The power <strong>and</strong> promise<br />

of population genomics: From genotyping to genome typing. Nature Reviews Genetics, 4:<br />

981-994.<br />

Sahin, K. <strong>and</strong> Gumusay, M. U., 2008. Service oriented architecture (SOA) based web services<br />

for geographic information systems. XXIst ISPRS Congress, Beijing, pp. 625-630.<br />

Service, R.F., 2006. The race for the 1000$ genome. Science, 311: 1544-1546.<br />

Tersoff, J., 2001. Nanotechnology: Less is more. Nature, 412: 135-136.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Road ecology: improving connectivity


S.R. Freitas et al. 2010. The effect of highways on native vegetation <strong>and</strong> reserve distribution in the State of São Paulo<br />

674<br />

The effect of highways on native vegetation <strong>and</strong> reserve distribution in<br />

the State of São Paulo, Brazil<br />

Simone R. Freitas 1* , Cláudia O. M. Sousa 2 & Jean Paul Metzger 2<br />

1 Universidade Federal do ABC (UFABC), Rua Santa Adélia, 166, 09210-170, Santo<br />

André, SP, Brazil<br />

2 Universidade de São Paulo (USP), Instituto de Biociências, Departamento de Ecologia,<br />

Rua do Matão, 321, travessa 14, 05508-900, São Paulo, SP, Brazil<br />

Abstract<br />

Highways affect the environment in different distances <strong>and</strong> intensities. This work aims to: 1)<br />

estimate areas which have been ecologically affected by highways in the whole state of São<br />

Paulo, for each type of vegetation, <strong>and</strong> in all nature reserves; <strong>and</strong>, 2) investigate the influence of<br />

highway distance on the native vegetation cover <strong>and</strong> on the reserve distribution. The area of<br />

study was the state of São Paulo (southeastern Brazil), where two biodiversity hotspots biomes<br />

occur: the Brazilian Atlantic <strong>Forest</strong> <strong>and</strong> the Brazilian Cerrado. About 10% of São Paulo has<br />

been ecologically affected by highways, being the dense ombrophylous forest <strong>and</strong> most nature<br />

reserves greatly impacted. Native vegetation <strong>and</strong> reserve areas have increased nearly logistically<br />

with the increase of highway distance. We suggest that in order to improve conservation <strong>and</strong><br />

restoration strategies, highways should be carefully considered, prioritizing remote areas.<br />

Keywords: road ecology, tropical forest, biodiversity conservation, tropical savanna, South<br />

America.<br />

1. Introduction<br />

Highways connect cities <strong>and</strong> facilitate transportation of products <strong>and</strong> services. They are a<br />

symbol of development for people living in remote areas (Pfaff et al. 2007). However, highways<br />

also affect air, soil, vegetation, wildlife (Forman et al. 2003). In tropical forest, the first effect of<br />

a highway construction is the forest fragmentation (Freitas et al. 2010), which cause edge effect<br />

<strong>and</strong> isolation of sensible species populations (Murcia 1995; Develey <strong>and</strong> Stouffer 2001;<br />

Fuentes-Montemayor et al. 2009; Laurance et al. 2009). Moreover, highways cause road kill,<br />

pollutants emission <strong>and</strong> facilitate fire events (Forman et al. 2003; Fahrig <strong>and</strong> Rytwinski 2009;<br />

Laurance et al. 2009). Some species show an road avoidance behavior, which reduces functional<br />

connectivity (McGregor et al. 2008; Fahrig <strong>and</strong> Rytwinski 2009), that is, the capacity of<br />

l<strong>and</strong>scape to facilitate biological flux (Taylor et al. 1993).<br />

The extension of the highway effects depends on the considered biotic or abiotic factor (Forman<br />

<strong>and</strong> Deblinger 1999). For instance, exotic plant species can reach up to 100 m from the road,<br />

whereas traffic noise can affect birds a hundred of meters far away from the road (Reijnen et al.<br />

1995; Forman <strong>and</strong> Deblinger 1999; Trombulak <strong>and</strong> Frissel 2000; Palomino <strong>and</strong> Carrascal 2007).<br />

Roads near to forest fragments can affect their species richness <strong>and</strong> composition (Hansen <strong>and</strong><br />

* Corresponding author. Tel.: +55 11 4437 8439<br />

Email address: simonerfreitas.ufabc@gmail.com<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.R. Freitas et al. 2010. The effect of highways on native vegetation <strong>and</strong> reserve distribution in the State of São Paulo<br />

675<br />

Clevenger 2005; Palomino <strong>and</strong> Carrascal 2007; Fahrig <strong>and</strong> Rytwinski 2009; Laurance et al.<br />

2009).<br />

Underst<strong>and</strong>ing the relationship between highways <strong>and</strong> native vegetation would improve<br />

methods to select priority areas for conservation <strong>and</strong> restoration <strong>and</strong> the effectivity of those new<br />

wildlife nature reserves. This work aims to: 1) estimate areas which have been ecologically<br />

affected by highways, in the whole state of São Paulo, for each type of vegetation, <strong>and</strong> in all<br />

nature reserves; <strong>and</strong>, 2) investigate the influence of highway distance on the native vegetation<br />

cover <strong>and</strong> on the reserve distribution.<br />

2. Methodology<br />

The study area was the State of São Paulo, in the southeastern of Brazil. In that State, there are<br />

two biomes considered as biodiversity hotspots for conservation: Atlantic <strong>Forest</strong> <strong>and</strong> Cerrado<br />

(Myers et al. 2000).<br />

We used the state’s road map, produced by DER (2008) <strong>and</strong> classified by road types: nonpaved<br />

roads, paved roads with two lanes, highways (paved roads with four lanes) <strong>and</strong><br />

expressways (paved roads with at least four lanes <strong>and</strong> high traffic). We also used a native<br />

vegetation map with the following vegetation types: Savanna <strong>and</strong> Semideciduous Seasonal<br />

<strong>Forest</strong> (both from Cerrado biome), <strong>and</strong> Serra do Mar Coastal <strong>Forest</strong>s, Araucaria Moist <strong>Forest</strong>s,<br />

Mangroves <strong>and</strong> Atlantic Coast Restingas (from Atlantic <strong>Forest</strong> biome).<br />

We included only the nature reserves classified as Full Protection in the Brazilian<br />

Environmental Legislation (SNUC) performing 62 nature reserves in the State of São Paulo.<br />

We estimated the areas which have been ecologically affected by roads in the whole State of<br />

São Paulo, in each type of vegetation <strong>and</strong> in the nature reserves following the methodology<br />

used by Forman (2000). The area ecologically affected by roads was evaluate using the<br />

distance where the sensible bird species are affected because they are negativelly affected by<br />

roads (Reijnen et al. 1995, Forman 2000, Develey e Stouffer 2001). Roads with high traffic<br />

flow were considered those with the higher ecological effect, thus the buffer width varied to<br />

road type (Forman 2000, Liu et al. 2008): non-paved roads have 200 m buffer width, paved<br />

roads 365 m, highways 810 m <strong>and</strong> expressways 1000 m.<br />

The relationship between native vegetation <strong>and</strong> road distance was evaluated using 10 noninclusive<br />

buffers: 0-50 m, 50-100 m, 100-250 m, 250-500 m, 500-750 m, 750-1000 m, 1000-<br />

1250 m, 1250-1500 m, 1500-1750 m, 1750-2000 m. In each buffer, we measured the total area<br />

of native vegetation.<br />

3. Result<br />

Road density in the São Paulo State was 0.145 km/km 2 . Paved roads were the most abundant<br />

(0.070 km/km 2 ), followed by non-paved roads (0.060 km/km 2 ), highways (0.010 km/km 2 ) <strong>and</strong><br />

expressways (0.006 km/km 2 ).<br />

More than 2,375,600 ha (10%) of the territory was affected ecologically by roads. The State of<br />

São Paulo was most affected by paved roads (4.7%), followed by non-paved roads (2.2%),<br />

highways (1.5%) <strong>and</strong> expressways (1.2%).<br />

About 5.4% of the native vegetation cover were affected by roads. The Atlantic <strong>Forest</strong> biome<br />

has about 5.3% of its cover affected by roads, whereas 5.9% of the Cerrado biome were affected<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.R. Freitas et al. 2010. The effect of highways on native vegetation <strong>and</strong> reserve distribution in the State of São Paulo<br />

676<br />

by them. Between the ecoregions (Olson et al. 2001), Serra do Mar coastal forests were the most<br />

affected by roads (51%). However, in relation to its cover, mangroves have more of its cover<br />

affected by roads (16%). About 55% of nature nature reserves were affected by roads. Five of<br />

them showed more than 30% of their territory affected by roads.<br />

There is more native vegetation cover as road distance increases (Figura 1). This relationship<br />

was nearly logistic for expressways (Figure 1).<br />

4. Discussion<br />

The State of São Paulo has a lower road density (0.15 km/km 2 ) than that considered as<br />

maximum for sustain populations of big predators (0.60 km/km 2 ; Forman <strong>and</strong> Alex<strong>and</strong>er 1998).<br />

However, those roads affect the distribution of native vegetation <strong>and</strong> threat the efficiency of<br />

nature reserves. Near roads could cause a significant increase of mortality rates for many<br />

wildlife populations, even overcoming hunting (Forman <strong>and</strong> Alex<strong>and</strong>er 1998).<br />

Almost 10% of State of São Paulo is ecologically affected by roads. Forman (2000) found a<br />

double proportion (about 20%) to United States of America. However, the State of São Paulo<br />

has a very lower road density than USA (0.41 km/km 2 ; Forman 2000), indicating that more<br />

native vegetation is under threat of roads in our study area. Even many states of USA have<br />

higher road densities than State of São Paulo (0.15 km/km 2 ): Misouri (1.89 km/km 2 ), Arkansas<br />

(1.23km/km 2 ) <strong>and</strong> Oklahoma (1.18 km/km 2 ; La Rue <strong>and</strong> Nielsen 2008).<br />

Laurance et al. (2009) state that non-paved roads represent a lower impact on vegetation <strong>and</strong><br />

wildlife in tropical forests than paved roads because they usually are inaccessible during raining<br />

season (summer). However, our study showed that non-paved <strong>and</strong> paved roads affect more than<br />

highways <strong>and</strong> expressways because they are more abundant <strong>and</strong> spread all over the territory,<br />

showing higher densities (0.06 km/km 2 <strong>and</strong> 0.07 km/km 2 respectivelly) than highways (0.01<br />

km/km 2 ) <strong>and</strong> expressways (0.01 km/km 2 ).<br />

As expected, the dominant vegetation type - Serra do Mar Coastal <strong>Forest</strong>s - was the most<br />

affected by roads. Serra do Mar Coastal <strong>Forest</strong>s is distributed along coastline as well as many of<br />

road network in Brazil, representing one of the most important connection axis for national<br />

transportation (north-south). However, in proportion mangroves are highly affected by roads for<br />

the same reason.<br />

Half of natural reserves is affected by roads. Five of them have more than 30% <strong>and</strong> three of<br />

them have more than 60% of their territory affected by roads, which represents a threat for their<br />

biodiversity. Roads facilitate hunter access <strong>and</strong> increase the probability of collision by vehicles<br />

(Laurance et al. 2009). Nature reserves near roads are probably under more conflicts with<br />

human population <strong>and</strong> more vulnerable to environmental degradation.<br />

There is more native vegetation as far as the road is. Roads act as atractor to l<strong>and</strong> use <strong>and</strong><br />

deforestation (Nagendra et al. 2003; Freitas et al. 2010). For instance, in the Amazon <strong>Forest</strong>,<br />

about 95% of deforestation occurred at most 50 km far from roads (Laurance et al. 2009).<br />

Expressways are not too densely distributed as other road types, however they show a nearly<br />

logistic relationship to native vegetation cover indicating a strong negative effect to the nearby<br />

native vegetation. Thus, road distance could be used as indicator to predict the distribution of<br />

native vegetation in the future. We suggest that in order to improve conservation <strong>and</strong> restoration<br />

strategies, the effect of highways should be carefully considered, prioritizing remote areas.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.R. Freitas et al. 2010. The effect of highways on native vegetation <strong>and</strong> reserve distribution in the State of São Paulo<br />

677<br />

References<br />

Develey, P.F. <strong>and</strong> Stouffer, P.C., 2001. Effects of roads on movements by understory birds in<br />

mixed-species flocks in central Amazonian Brazil. Conservation Biology, 15: 1416-1422.<br />

Fahrig, L. <strong>and</strong> Rytwinski, T., 2009. Effects of roads on animal abundance: an empirical review<br />

<strong>and</strong> synthesis. Ecology <strong>and</strong> Society, 14: 21.<br />

Forman, R.T.T. <strong>and</strong> Alex<strong>and</strong>er, L.E., 1998. Roads <strong>and</strong> their major ecological effects. Annual<br />

Reviews in Ecology & Systematics, 29: 207-231.<br />

Forman, R.T.T. <strong>and</strong> Deblinger, R.D., 2000. The ecological road-effect zone of a Massachusetts<br />

(U.S.A.) suburban highway. Conservation Biology, 14: 36-46.<br />

Forman, R.T.T., 2000. Estimate of the area affected ecologically by the road system in the<br />

United States. Conservation Biology, 14: 31-35.<br />

Forman, R.T.T., Sperling, D., Bissonette, J.A., Clevenger, A.P., Cutshall, C.D., Dale, V.H.,<br />

Fahrig, L., France, R., Goldman, C.R., Heanue, K., Jones, J.A., Swanson, F.J., Turrentine,<br />

T. <strong>and</strong> Winter, T.C., 2003. Road ecology: science <strong>and</strong> solutions. Isl<strong>and</strong> Press,<br />

Washington, DC, 481 p.<br />

Freitas, S.R., Hawbaker, T.J. <strong>and</strong> Metzger, J.P., 2010. Effects of roads, topography, <strong>and</strong> l<strong>and</strong><br />

use on forest cover dynamics in the Brazilian Atlantic <strong>Forest</strong>. <strong>Forest</strong> Ecology <strong>and</strong><br />

Management, 259: 410-417.<br />

Fuentes-Montemayor, E., Cuarón, A.D., Vázquez-Domínguez, E., Benítez-Malvido, J.,<br />

Valenzuela-Galván, D. <strong>and</strong> Andresen, E., 2009. Living on the edge: roads <strong>and</strong> edge<br />

effects on small mammal populations. Journal of Animal Ecology, 78: 857-865.<br />

Hansen, M.J. <strong>and</strong> Clevenger, A.P., 2005. The influence of disturbance <strong>and</strong> habitat on the<br />

presence of non-native plant species along transport corridors. Biological Conservation,<br />

125: 249-259.<br />

LaRue, M.A. <strong>and</strong> Nielsen, C.K., 2008. Modelling potential dispersal corridors for cougars in<br />

midwestern North America using least-cost path methods. Ecological Modelling, 212:<br />

372-381.<br />

Laurance, W.F., Goosem, M. <strong>and</strong> Laurance, S.G.W., 2009. Impacts of roads <strong>and</strong> linear clearings<br />

on tropical forests. Trends in Ecology <strong>and</strong> Evolution, 24: 659-669.<br />

Liu, S.L., Cui, B.S., Dong, S.K., Yang, Z.F., Yang, M. <strong>and</strong> Holt, K., 2008. Evaluating the<br />

influence of road networks on l<strong>and</strong>scape <strong>and</strong> regional ecological risk - A case study in<br />

Lancang River Valley of Southwest China. Ecological Engineering, 34: 91-99.<br />

McGregor, R.L., Bender, D.J. <strong>and</strong> Fahrig, L., 2008. Do small mammals avoid roads because of<br />

the traffic Journal of Applied Ecology, 45: 117-123.<br />

Murcia, C., 1995. Edge effects in fragmented forests: implications for conservation. Trends in<br />

Ecology <strong>and</strong> Evolution, 10: 58-62.<br />

Nagendra, H., Southworth, J. <strong>and</strong> Tucker, C., 2003. Accessibility as a determinant of l<strong>and</strong>scape<br />

transformation in western Honduras: linking pattern <strong>and</strong> process. L<strong>and</strong>scape Ecology, 18:<br />

141-158.<br />

Palomino, D. <strong>and</strong> Carrascal, L.M., 2007. Threshold distances to nearby cities <strong>and</strong> roads<br />

influence the bird community of a mosaic l<strong>and</strong>scape. Biological Conservation, 140: 100-<br />

109.<br />

Pfaff, A., Robalino, J., Walker, R., Aldrich, S., Caldas, M., Reis, E., Perz, S., Bohrer, C., Arima,<br />

E., Laurance, W. <strong>and</strong> Kirby, K., 2007. Road investments, spatial spillovers, <strong>and</strong><br />

deforestation in the Brazilian Amazon. Journal of Regional Science, 47: 109-123.<br />

Reijnen, R., Foppen, R., Ter Braak, C. <strong>and</strong> Thissen, J., 1995. The effects of car traffic on<br />

breeding bird populations in woodl<strong>and</strong>. III. Reduction of density in relation to the<br />

proximity of main roads. Journal of Applied Ecology, 32: 187-202.<br />

Taylor, P.D., Fahrig, L., Henein, K. <strong>and</strong> Merriam, G., 1993. Connectivity is a vital element of<br />

l<strong>and</strong>scape structure. Oikos, 68: 571-573.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S.R. Freitas et al. 2010. The effect of highways on native vegetation <strong>and</strong> reserve distribution in the State of São Paulo<br />

678<br />

Trombulak, S.C. <strong>and</strong> Frissell, C.A., 2000. Review of ecological effects of roads on terrestrial<br />

<strong>and</strong> aquatic communities. Conservation Biology,14: 18-30.<br />

18<br />

native vegetation cover / buffer area * 100<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

50 100 250 500 750 1000 1250 1500 1750 2000<br />

distance from road (m)<br />

expressway highway paved road unpaved road express_logist<br />

Figure 1: Native vegetation cover (%) in different distances from roads for each road type. Native<br />

vegetation cover increases nearly logistically (express_logist) in function to distance of expressways.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Terrones et al. 2010. Detecting vulnerable spots for ecological connectivity caused by minor road network<br />

679<br />

Detecting vulnerable spots for ecological connectivity caused by minor<br />

road network in Alicante, Spain<br />

Beatriz Terrones 1,2* , Sara Michelle Catalán 1,3 , Aydeé Sanjinés 1,3 , Encarnación Rico-<br />

Guzmán 4 & Andreu Bonet 1,4<br />

1 Department of Ecology. University of Alicante. Spain<br />

2 Font Roja NATURA UA' Scientific Station. University of Alicante. Spain<br />

3 Mediterranean Agronomic Institute of Zaragoza. CIHEAM, Spain<br />

4 "Ramón Margalef" Multidisciplinary Institute for Environmental Research. University<br />

of Alicante, Spain<br />

Abstract<br />

Applying l<strong>and</strong> conservation planning allows to maintain <strong>and</strong> restore the ecological connections<br />

among natural remnants <strong>and</strong> protected areas in the territory. Roads increase the problem of<br />

habitat fragmentation, breaking large habitat areas into smaller ones, creating isolated habitat<br />

patches <strong>and</strong> decreasing connectivity in the territory.<br />

Functional connectivity in Alicante’s Province was analyzed using least-cost models. Applied<br />

connectivity models for the study of ecological processes <strong>and</strong> animal movements is a tool of<br />

great utility for l<strong>and</strong>scape planning <strong>and</strong> Protected Areas management. Connectivity models are<br />

based in the creation of a friction surface that indicates the relative cost of moving target species<br />

across the l<strong>and</strong>scape. Ecological corridors between the main protected areas <strong>and</strong> conflictive<br />

points that intersect with minor road network were identified.<br />

Keywords: ecological connectivity, l<strong>and</strong>scape model, cost-distance, road ecology.<br />

1. Introduction<br />

L<strong>and</strong> use changes in l<strong>and</strong>scape are the main cause of fragmentation processes of natural habitats<br />

<strong>and</strong> populations of wild organisms, <strong>and</strong> it is recognized as one of the most important threats to<br />

the biodiversity maintenance (Harris 1984; Wilson 1988; Saunders <strong>and</strong> Hobbs 1991;<br />

McCullough 1996; Pickett et al. 1997; Fielder <strong>and</strong> Kareiva 1998; Fahrig 2003). Linear<br />

infrastructures like roads <strong>and</strong> railway networks are important barriers that limit movement <strong>and</strong><br />

dispersion of terrestrial vertebrates, increasing the problem of habitat fragmentation <strong>and</strong> creating<br />

isolated habitat patches in the territory. The presence of a road may modify animal’s behaviour<br />

<strong>and</strong> alter patterns of animal movement (Trombulak 2000).<br />

It is fundamental for wildlife survival to maintain <strong>and</strong> restore the ecological connections among<br />

natural remnants <strong>and</strong> Protected Areas in the territory. The concept of ecological connectivity is<br />

shown as an essential element in the l<strong>and</strong>-use planning. L<strong>and</strong>scape connectivity could be<br />

defined as the degree to which the l<strong>and</strong>scape facilitates or impedes movement of organisms<br />

among resource patches (Taylor et al. 1993). This definition considers that connectivity is<br />

* Corresponding author. Tel.: 965 903 732 - Fax: 965 909 832<br />

Email address: beatriz.terrones@ua.es<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Terrones et al. 2010. Detecting vulnerable spots for ecological connectivity caused by minor road network<br />

680<br />

species <strong>and</strong> l<strong>and</strong>scape specific, because it depends not only on characteristics of the l<strong>and</strong>scape,<br />

but also on aspects of the mobility of the organism (Tischendorf <strong>and</strong> Fahrig 2000).<br />

Application of connectivity models for the study of ecological processes <strong>and</strong> species movement<br />

is a useful tool in conservation planning (Adriansen et al. 2003; Nikolakaki 2004; Marulli <strong>and</strong><br />

Mallarach, 2005; Driezen et al. 2007) <strong>and</strong> it can be an interesting tool for predicting the effect<br />

of changes in the l<strong>and</strong>scape on connectivity in a quantitative way (Adriansen et al. 2003). These<br />

models produce a map of the relative cost of reaching particular areas in a l<strong>and</strong>scape from a<br />

source.<br />

This work presents a methodological spatial approach based on Geographical Information<br />

System (GIS) that (1) enables a connectivity diagnose of terrestrial l<strong>and</strong>scape ecosystems <strong>and</strong><br />

(2) makes possible the identification of vulnerable spots that have a critical importance for<br />

ecological connectivity. These results will provide useful guidelines for l<strong>and</strong>scape conservation<br />

planning at local level (Municipalities <strong>and</strong> Local Administrations).<br />

2. Methodology<br />

2.1 Habitat maps<br />

We created a l<strong>and</strong> use map for Alicante’s Province compiling information about l<strong>and</strong> use <strong>and</strong><br />

vegetation cover from different sources. The original map was the Spanish <strong>Forest</strong> Map<br />

(MFE50), which was reclassified. We used CORINE l<strong>and</strong> cover map, road <strong>and</strong> railway network<br />

<strong>and</strong> urban areas cartography to complete it. The final map recognized 32 l<strong>and</strong>-use types, of<br />

which 7 are linear structures. In order to improve a finer analysis resolution <strong>and</strong> to incorporate<br />

barriers <strong>and</strong> connection spots at l<strong>and</strong>scape level, we also created derived cartography, as road<br />

network impact <strong>and</strong> urban areas impact maps (decreasing connectivity), <strong>and</strong> hydrological<br />

network map <strong>and</strong> tunnels <strong>and</strong> bridges map (improving connectivity). These maps were<br />

converted to raster format, with a cell resolution of 20x20 meters.<br />

2.2 Least-cost analysis<br />

Connectivity models are based on the creation of a friction surface that indicates the relative<br />

cost of moving for target species across the l<strong>and</strong>scape (Figure 1). We used raster format, from<br />

ArcGIS software, with a cell resolution of 20x20 meters.<br />

Figure 1: Methodology of the process followed.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Terrones et al. 2010. Detecting vulnerable spots for ecological connectivity caused by minor road network<br />

681<br />

We defined target species as forest mammals present in Alicante’s Province, such as the stone<br />

marten (Martes foina Erxleben, 1777), the red fox (Vulpes vulpes Linnaeus, 1758), the wildcat<br />

(Felis silvestris Schreber, 1777) <strong>and</strong> the wild boar (Sus scrofa Linnaeus, 1758). These species<br />

use mainly riversides <strong>and</strong> ravines to move along the l<strong>and</strong>scape, while urban areas <strong>and</strong> road<br />

network are the main barriers to their movements. We identified the core areas of target species<br />

as dense forested areas inside Sites of Community Interest (SCI) of Natura 2000, with a<br />

minimum area of 20 ha (McDonald <strong>and</strong> Barret 2008).<br />

We determined the main l<strong>and</strong>scape factors that have more influence on the resistance to forest<br />

mammals movement, creating a final friction map on GIS with a relative set of resistance values<br />

for l<strong>and</strong> use covers <strong>and</strong> the other cartographic information layers (Figure 2), based on previous<br />

works (Sastre et al. 2002; Campeny et al. 2005; Gurrrutxaga 2005).<br />

Figure 2: Friction map for Alicante’s Province, with a relative scale from 0.6 to 701.5.<br />

With this information, we created a cost-distance surface <strong>and</strong> calculated the least-cost route<br />

between natural areas. Finally, we carried out an identification of l<strong>and</strong>scape linkages <strong>and</strong><br />

ecological corridors <strong>and</strong> their intersection with the minor road network, in order to propose<br />

restoration activities to improve connectivity.<br />

3. Results<br />

We obtained a final map of the cost distance from habitat sources for target species (Figure 3).<br />

With this map, we were able to analyze the relative cost of reaching points in the l<strong>and</strong>scape<br />

from a determined source. In this case, we analyzed the problem with connectivity between the<br />

Natura 2000 areas, <strong>and</strong> their intersections with the road network.<br />

Results showed that core areas in the North of the Province were more connected than core<br />

areas in the South, caused by the presence of more natural areas <strong>and</strong> forests in the North.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Terrones et al. 2010. Detecting vulnerable spots for ecological connectivity caused by minor road network<br />

682<br />

Figure 3: Cost distance map for Alicante’s Province, in a relative scale from 0 to 100.<br />

The cost distance values for the Alicante Province surface are showed in Table 1. The 20% of<br />

the area of Alicante’s Province was classified with a very high level of connection <strong>and</strong> only the<br />

9% of area was disconnected.<br />

Table 1: Cost values in a categorical relative scale (from 0 to 100) of the Alicante’s Province.<br />

Category Cost values Total Area (Km 2 ) %<br />

Very high connectivity 0 – 1 1211.23 20.84<br />

High connectivity 1 – 5 1975.92 33.99<br />

Medium connectivity 5 – 10 1335.85 22.98<br />

Low connectivity 10 – 20 765.31 13.17<br />

No connectivity 20 – 100 524.28 9.02<br />

Based on the cost distance map, we selected the areas with the lowest values in travel cost<br />

between the Natura 2000 areas, <strong>and</strong> then we proposed the main corridors between protected<br />

areas. We detected also the intersections between these corridors <strong>and</strong> the road network. Then,<br />

we used the cartography of transversal structures along the roads (tunnels <strong>and</strong> bridges) as<br />

connection spots, <strong>and</strong> searched for the conflictive areas for connectivity.<br />

A good example of the application of this work, it can be observed in the North area of the<br />

Province, in the Maigmó i Serres de la Foia de Castalla SCI in a fragmented l<strong>and</strong>scape (Figure<br />

4). This Protected Area is surrounding by other natural areas, which are core areas for the target<br />

species. CV-80 <strong>and</strong> A-31 are two highways that limited connections by North <strong>and</strong> West. In<br />

these highways there are several transversal structures that could be adapted for improving<br />

connectivity. In the East area there are two minor roads (CV-802 <strong>and</strong> CV- 803) that are limiting<br />

connectivity, but the cost values are lower than in the other areas.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Terrones et al. 2010. Detecting vulnerable spots for ecological connectivity caused by minor road network<br />

683<br />

Figure 4: Detail of cost surface map of the Maigmó <strong>and</strong> Serres de la Foia de Castalla SCI, with the road<br />

network in black, bridges as white points, <strong>and</strong> arrows defining the main areas for wildlife movements.<br />

4. Discussion<br />

The methodology we use for evaluating ecological connectivity at a regional scale allows<br />

obtaining quick assessments, which can be very effective for l<strong>and</strong>-use planning, <strong>and</strong> it is also a<br />

flexible research tool. This methodology has a lot of potentials to study connectivity problems<br />

with a wide range of applications (Adriansen et al. 2003; Pinto <strong>and</strong> Keitt 2009).<br />

This analysis allows to identify important points for ecological connectivity such as vulnerable<br />

spots <strong>and</strong> areas with high value for ecological connectivity. The generated maps will provide<br />

useful guidelines for l<strong>and</strong>scape conservation planning at local level (Municipalities <strong>and</strong> Local<br />

Administrations).<br />

Connectivity between Natura 2000 Network sites in Alicante’s Province is heterogeneous. The<br />

North of Province shows the highest values of cost distance, meanwhile there are many areas<br />

disconnected in the South. The 22% of the Province surface have low or no connectivity. These<br />

areas would need the target of restoration <strong>and</strong> permeability actions, but also areas with high<br />

connectivity in order to solve existing conflicts between ecological <strong>and</strong> road network.<br />

Roads <strong>and</strong> urban areas appear as one of the most important problems for connectivity in<br />

Alicante’s Province. The intersection of proposed corridors based on cost distance surface maps,<br />

with the road network cartography would be a good tool for selecting conflict spots in order to<br />

propose restoration activities to improve connectivity.<br />

Acknowledgements<br />

This research was supported by Alicante Natura S.A., Diputación de Alicante <strong>and</strong> the project<br />

“Balances hídricos y recarga de acuíferos en sistemas mediterráneos: Comparación de<br />

escenarios de cambio de usos del suelo y en un contexto de cambio climático (BAHIRA)”<br />

(CGL2008-03649/BTE).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


B. Terrones et al. 2010. Detecting vulnerable spots for ecological connectivity caused by minor road network<br />

684<br />

References<br />

Adriansen, F.; Chardon, J. P.; DeBlust, G.; Swinnen, E.; Villalba, S.; Gulinck, G. And<br />

Matthysen, E., 2003. The application of ‘least-cost’ modeling as a functional l<strong>and</strong>scape<br />

model. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 64: 233–247.<br />

Campeny, R.; Brotons, L.; Pla, M. <strong>and</strong> Rosell, C., 2005. Obtenció de mapes de resistencia a la<br />

dispersió de mamífers i avaluació de l’interès de conservació del territori dins el mòdul<br />

de fauna i connectivitat del SIXTELL. Desembre 2005. Diputació Barcelona. Minuartia.<br />

Informe Inédito.<br />

Driezen, K.; Adriaensen, F.; Rondinini, C.; Doncaster, C.P. <strong>and</strong> Matthysen, E., 2007. Evaluating<br />

least-cost model predictions with empirical dispersal data: A case-study using<br />

radiotracking data of hedgehogs (Erinaceus europaeus). Ecological Modelling, 209: 314<br />

– 322.<br />

Fahrig, L., 2003. Effects of habitat fragmentation on biodiversity. Annual Review of Ecology,<br />

Evolution <strong>and</strong> Systematics, 34: 487-515.<br />

Fielder, P.L. <strong>and</strong> Kareiva, P.M. (eds.), 1998. Conservation Biology for the Coming Decade.<br />

Chapman & Hall, New York.<br />

Gurrutxaga, M., 2005. Red de Corredores Ecológicos de la Comunidad Autónoma de Euskadi.<br />

Informe inédito de IKT para la Dirección de Biodiversidad del Gobierno Vasco. 150 pp.<br />

Harris, L.D., 1984. The Fragmented <strong>Forest</strong>: Isl<strong>and</strong> Biogeographic Theory <strong>and</strong> the Preservation<br />

of Biotic Diversity. University of Chicago Press, Chicago.<br />

Marulli, J. <strong>and</strong> Mallarach, J.M., 2005. A GIS methodology for assensing ecological<br />

connectivity: application to the Barcelona Metropolitan Area. L<strong>and</strong>scape <strong>and</strong> Urban<br />

Planning, 71: 243-262.<br />

McCullough, D.R. (ed.), 1996. Metapopulations <strong>and</strong> Wildlife Conservation. Isl<strong>and</strong> Press,<br />

Washington, DC.<br />

McDonald, D. <strong>and</strong> Barret, P., 2008. Guía de campo de los mamíferos de España y de Europa.<br />

Ed.Omega, Barcelona.<br />

Nikolakaki, P., 2004. A GIS site-selection process for habitat creation: Estimating connectivity<br />

of habitat patches. L<strong>and</strong>scape <strong>and</strong> Urban Planning, 68:77-94.<br />

Pickett, S.TA.; Ostfeld, R.S.; Shachak, M. <strong>and</strong> Likens, G.E. (eds.), 1997. The Ecological Basis<br />

of Conservation: Heterogeneity, ecosystem, <strong>and</strong> Biodiversity. Chapman & Hall, New<br />

York.<br />

Pinto, N. <strong>and</strong> Keitt, T.H., 2009. Beyond the least-cost path: evaluating corridor redundancy<br />

using a graph-theoretic approach. L<strong>and</strong>scape Ecology, 24: 253-266.<br />

Sastre, P.; de Lucio, J.V. <strong>and</strong> Martínez, C., 2002. Modelos de conectividad del paisaje a<br />

distintas escalas. Ejemplos de aplicación en la Comunidad de Madrid. Ecosistemas<br />

2002/2.<br />

Saunders, D. <strong>and</strong> R. Hobbs (eds.), 1991. Nature Conservation 2: The Role of Corridors. Surrey<br />

Beatty, Chipping Norton, Australia.<br />

Taylor, P.D.; Fahrig, L.; Henein, K. <strong>and</strong> Merriam, G., 1993. Connectivity is a vital element of<br />

l<strong>and</strong>scape structure. Oikos 68:571–573.<br />

Tischendorf, L. <strong>and</strong> Fahrig, L., 2000. On the usage <strong>and</strong> measurement of l<strong>and</strong>scape connectivity.<br />

Oikos, 90: 7-19.<br />

Trombulak, S.C. <strong>and</strong> Frissell, C. A., 2000. Review of ecological effects of roads on terrestrial<br />

<strong>and</strong> aquatic communities. Conservation Biology, 14(1): 18-30.<br />

Wilson, E.O., 1988. The current status of biological diversity. In: Wilson E.O. (ed.),<br />

Biodiversity. National Academic Press, Washington, DC.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A l<strong>and</strong>scape approach to sustainable forest management


M. Kolström et al. 2010. Is it possible to combine adaptation to climate change <strong>and</strong> maintaining of forest biodiversity<br />

686<br />

Is it possible to combine adaptation to climate change <strong>and</strong> maintaining of<br />

forest biodiversity<br />

Marja Kolström * , Terhi Vilén & Marcus Lindner<br />

European <strong>Forest</strong> Institute, Finl<strong>and</strong><br />

Abstract<br />

An EU-level review of the climate change adaptation measures in forestry shows that there are<br />

adaptation measures that support maintaining biodiversity, but also adaptation measures that have<br />

the potential to decrease the level of biodiversity. A choice of the adaptation measure might thus<br />

involve trade-offs between efficient adaptation <strong>and</strong> maintaining biodiversity at the st<strong>and</strong> level. The<br />

l<strong>and</strong>scape approach allows building a combination of highly adaptive st<strong>and</strong>s with simultaneous high<br />

level of biodiversity <strong>and</strong> st<strong>and</strong>s where adaptation measures do not support biodiversity. This is<br />

however a complex task <strong>and</strong> gets even more complicated since also economic <strong>and</strong> social<br />

st<strong>and</strong>points have to be taken into account. A successful realisation is possible only if all policy<br />

makers at different levels, affected stakeholder groups, forest owners <strong>and</strong> forest workers are aware<br />

what measures are suitable <strong>and</strong> why they are used.<br />

Keywords: European forests, climate change, adaptation measures, l<strong>and</strong>scape approach,<br />

biodiversity<br />

1. Introduction<br />

Rising levels of greenhouse gases are already changing the climate. According to the IPCC WGI<br />

Fourth Assessment Report, the global average temperature has increased by about 0.76 ºC from<br />

1850 to 2005, <strong>and</strong> a further increase in temperatures of 1.4°C to 5.8°C by 2100 is projected. Thus,<br />

merely mitigation of the climate change is not enough but also adaptation of forest management is<br />

essential to avoid negative impacts for the forestry sector <strong>and</strong> to ensure the continuation of the<br />

mitigation effect that forests have. In addition to changes in mean climate variables, European<br />

forests will have to adapt also to increased climate variability; there is expected to be greater risk of<br />

extreme weather events like prolonged drought, storms <strong>and</strong> floods (Maracchi et al. 2005, Salinger et<br />

al. 2005).<br />

European forests are diverse, with each region featuring different tree species, ecological<br />

conditions, management goals, <strong>and</strong> required goods <strong>and</strong> services by the society. Thus, the adaptive<br />

capacity of the European forests differs <strong>and</strong> also the climate change adaptation strategies should be<br />

developed in different ways. Main goals of the adaptive management include maintaining the wood<br />

production (boreal region), minimizing the impacts of disturbances (temperate <strong>and</strong> Mediterranean<br />

region) <strong>and</strong> ensuring the ecosystem services (Mediterranean region) (Lindner et al. 2008).<br />

* Corresponding author. Tel.: +358 10 773 4334 - Fax: +358 10 773 4377<br />

Email address: marja.kolstrom@efi.int<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


M. Kolström et al. 2010. Is it possible to combine adaptation to climate change <strong>and</strong> maintaining of forest biodiversity<br />

687<br />

Furthermore, also biodiversity has to be taken into account when planning the adaptive management<br />

of the forests. In addition of having intrinsic value, biodiversity can also play an important role in<br />

enhancing the adaptive capacity of the forest ecosystems, i.e. increasing the ecosystem capacity to<br />

recover <strong>and</strong> adapt to the impacts of climate change (Secretariat of the Convention on Biological<br />

Diversity 2003).<br />

Existing knowledge of the observed <strong>and</strong> projected impacts of the climate change on European<br />

forests, <strong>and</strong> options for forests <strong>and</strong> forestry to adapt to climate change were reviewed on a study<br />

commissioned by the Directorate General for Agriculture <strong>and</strong> Rural Development of the European<br />

Commission (Lindner et al. 2008). This EU-level study shows that there are adaptation measures<br />

that support maintaining biodiversity, but also adaptation measures that have the potential to<br />

decrease the level of biodiversity. A choice of the adaptation measure might thus involve trade-offs<br />

between efficient adaptation <strong>and</strong> maintaining biodiversity at the st<strong>and</strong> level. Consequently, planning<br />

<strong>and</strong> management of the adaptation measures should be carried out at the l<strong>and</strong>scape level. The aim of<br />

this study is to review adaptation options in forestry <strong>and</strong> their effect on biodiversity <strong>and</strong> then assess<br />

how l<strong>and</strong>scape approach might be a solution to combine adaptation to climate change <strong>and</strong><br />

maintaining of forest biodiversity.<br />

2. Adaptation options in forestry <strong>and</strong> interaction with biodiversity<br />

Species or st<strong>and</strong> composition of forest can be manipulated in a regeneration phase. Because the<br />

genetic composition of the populations needs to be enhanced to cope with environmental changes,<br />

diversity should be set at higher levels than under st<strong>and</strong>ard regeneration conditions, i.e. without<br />

environmental change. Thus, in the regeneration phase, a highly recommended option to secure the<br />

adaptive response of established regeneration is to raise the level of genetic diversity within the<br />

seedling population, either by natural or artificial mediated means. In order to raise levels of genetic<br />

diversity, seedlings coming from different seed st<strong>and</strong>s of the same provenance region can be mixed<br />

at the nursery stage. This introducing of new reproductive material should however be seen as<br />

complementing local seed resources, not replacing local material (Lindner et al. 2008).<br />

Adaptation measures in early st<strong>and</strong> development stages i.e. tending <strong>and</strong> pre-commercial thinning)<br />

should support mixed st<strong>and</strong>s of suitable, adapted tree species, inter alia to distribute risk via<br />

diversification (Spiecker 2003). Silvicultural adaptation actions aiming at the establishment of<br />

ecosystems with highly diverse tree composition <strong>and</strong> ground vegetation will be inevitable in pests<br />

<strong>and</strong> disease risk management. Because of increasingly dry periods, fire risk is expected to increase<br />

in the already fire‐prone Mediterranean zone, but also in the boreal <strong>and</strong> central European regions.<br />

Current fire prevention policies need to be adjusted to cope with longer <strong>and</strong> severe fire seasons <strong>and</strong><br />

larger areas exposed to fire risk, especially in the Mediterranean region (Moreno 2005). Adaptation<br />

measures include, for example, modification of forest structure (e.g. tree spacing <strong>and</strong> density,<br />

regulation of age class structure) <strong>and</strong> fuel management, like thinning, pruning <strong>and</strong> biomass<br />

removals. To decrease risk of pests or fire, removal of forest residues, wind throws clearings,<br />

sanitation felling <strong>and</strong> removing st<strong>and</strong>ing dead trees <strong>and</strong> coarse woody debris on the forest floor are<br />

recommended. However, mature trees <strong>and</strong> decaying wood, for instance, are elements favouring<br />

diversity in forest ecosystems.<br />

With term harvesting we refer to production-target oriented cutting of mature trees, ranging from<br />

single individuals (e.g. selection cutting) to larger spatial scales (e.g. clear cut). Small scale<br />

harvesting interventions <strong>and</strong> the promotion of harvesting systems which support natural<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


M. Kolström et al. 2010. Is it possible to combine adaptation to climate change <strong>and</strong> maintaining of forest biodiversity<br />

688<br />

regeneration of suitable species increase spatial heterogeneity (e.g. Spiecker 2003). A dense forest<br />

road network is the prerequisite to small-scale, structurally diverse thinning <strong>and</strong> harvesting practices<br />

under adaptive management. In complex terrain such as mountain forests, the vital forest functions<br />

are depending on such measures (Brang et al. 2006; Woltjer et al. 2008). Moreover, infrastructure<br />

supports the mitigation of large scale disturbance impacts. This however leads to fragmentation of<br />

ecosystems <strong>and</strong> has adverse effects to the biodiversity.<br />

Reducing forest fragmentation by establishing connecting corridors between densely forested<br />

regions contributes to increasing the natural adaptive capacity. Establishing a network of corridors<br />

<strong>and</strong> stepping stones connecting protected areas has been suggested as an adaptive strategy to assist<br />

species as they migrate in response to climate change (Halpin 1997; Harrington et al. 2001).<br />

Corridors should be built to aid in the migration of species to other protected areas when an extreme<br />

event has occurred (Bridgewater <strong>and</strong> Woodin 1990).<br />

<strong>Forest</strong> management planning describes the processes of problem identification, development of<br />

alternatives <strong>and</strong> selection of alternatives, embedded in an adaptive management cycle (cf. Rauscher<br />

1999). One significant selection is the length of rotation, i.e. the period of years between when a<br />

forest st<strong>and</strong> is established <strong>and</strong> when it receives its final harvest. A shortening of rotation length is an<br />

adaptation option in regions with increasing forest growth, like in mountainous or boreal<br />

environments. The reduced rotation length will also lower the risk of financial losses due to<br />

calamities <strong>and</strong> counteract the reduced management flexibilities induced by excessive levels of<br />

sanitation felling (e.g. Wermelinger 2004). However, structural changes of lower rotation lengths<br />

need to be considered with regard to other forest services, like biodiversity via loss of large<br />

diameter trees.<br />

3. L<strong>and</strong>scape approach may be a solution<br />

L<strong>and</strong>scape level approach would allow building a combination of highly adaptive st<strong>and</strong>s with<br />

simultaneous high level biodiversity <strong>and</strong> st<strong>and</strong>s where adaptation measures do not support<br />

biodiversity. This is however a complex task <strong>and</strong> gets even more complicated since also economic<br />

<strong>and</strong> social st<strong>and</strong>points have to be taken into account. The measure might be relevant from<br />

ecological point of view, but not in economic or social point of view.<br />

Since disturbance agents will gain more importance under the climate change, l<strong>and</strong>scape level forest<br />

planning, balancing st<strong>and</strong> level management goals <strong>and</strong> integrated forest protection measures, is of<br />

increasing importance. A key approach in risk management is diversification of tree species<br />

mixtures <strong>and</strong> management approaches between neighbouring forest st<strong>and</strong>s or within a forestry<br />

district to increase adaptive capacity (Lindner 2000; Bodin <strong>and</strong> Wiman 2007; Lindner 2007).<br />

In developing alternative management strategies, forest planning aggregates st<strong>and</strong> scale<br />

management options <strong>and</strong> aims for concerted application at the l<strong>and</strong>scape level. Promotion of<br />

diversity in l<strong>and</strong>scape structure is seen as a strategy to promote resilience of forest ecosystems. At<br />

larger geographical scales of management units <strong>and</strong> forest l<strong>and</strong>scapes, various strategies can be<br />

combined. For example, preferring natural regeneration in mixed forests with long rotation cycles is<br />

incompatible with planting productive genotypes managed in short rotation cycles.<br />

Identifying suitable l<strong>and</strong>scape level management strategies can be facilitated with decision support<br />

systems which include necessary simulation models. Integrated ecosystem models, for instance,<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


M. Kolström et al. 2010. Is it possible to combine adaptation to climate change <strong>and</strong> maintaining of forest biodiversity<br />

689<br />

offer means to consistently project effects of adapted management on a variety of sustainable forest<br />

management indicators. The development of integrated, model-based decision support tools relying<br />

on multi-criteria analysis as a means to integrate stakeholders could offer means to improve the<br />

science – policy –practice interface.<br />

A benefit of the l<strong>and</strong>scape approach is that it includes also socio-economic aspects. The<br />

socioeconomic adaptive capacity of some region can be improved by investing into infrastructure,<br />

research <strong>and</strong> increasing awareness of suitable adaptation measures. One example of a link between<br />

ecological <strong>and</strong> social aspects is conservation areas; two or more protected areas can be linked to one<br />

another with corridors, which aim to help protect the biodiversity. Biological corridors are sections<br />

of l<strong>and</strong> managed via a combination of voluntary agreements or compensation packages with the<br />

owners of the l<strong>and</strong> (McNeely 1994).<br />

4. Discussion<br />

There are quite many examples of adaptive management measures to climate change which are<br />

supporting biodiversity of forests. But there are also measures which are harmful from biodiversity<br />

point of view. Thus, is it possible to combine adaptation to climate change <strong>and</strong> maintaining of forest<br />

biodiversity The l<strong>and</strong>scape approach is a good solution to this complex situation. If a measure<br />

provides win-win –situation for adaptation <strong>and</strong> biodiversity, it should be certainly used. Otherwise<br />

we should choose our object either adaptation or biodiversity for a st<strong>and</strong>, <strong>and</strong> at l<strong>and</strong>scape level<br />

have a good balance of the measures to achieve good level of biodiversity as well as adaptation.<br />

A successful realisation of adaptation measures is possible only if all policy makers at different<br />

levels, affected stakeholder groups, forest owners <strong>and</strong> forest workers are aware what measures are<br />

suitable <strong>and</strong> why they are used. It is of utmost importance to disseminate the knowledge on suitable<br />

adaptation measures <strong>and</strong> their impacts on biodiversity to all stakeholder groups, who need to<br />

implement the measures on the ground. For example, risk management in forestry can be promoted<br />

by educational efforts (Zebisch et al. 2005). Appropriate tools <strong>and</strong> systems are also necessary to<br />

support dissemination.<br />

References<br />

Bodin, P. <strong>and</strong> Wiman, B.L.B., 2007. The usefulness of stability concepts in forest management<br />

when coping with increasing climate uncertainties. <strong>Forest</strong> Ecology <strong>and</strong> Management, 242:<br />

541-552.<br />

Brang, P., Schönenberger, W., Frehner, M., Schwitter, R., Thormann, J. <strong>and</strong> Wasser, B., 2006.<br />

Management of protection forests in the European Alps: An overview. <strong>Forest</strong> Snow <strong>and</strong><br />

L<strong>and</strong>scape Research, 80: 23-44.<br />

Bridgewater, P. <strong>and</strong> Woodin, S.J., 1990. <strong>Global</strong> warming <strong>and</strong> nature conservation. L<strong>and</strong> Use Policy,<br />

7: 165-168.<br />

Halpin, P.N., 1997. <strong>Global</strong> climate change <strong>and</strong> natural-area protection: Management responses <strong>and</strong><br />

research directions. Ecological Applications, 7: 828-843.<br />

Harrington, R., Fleming, R.A. <strong>and</strong> Woiwod, I.P., 2001. Climate change impacts on insect<br />

management <strong>and</strong> conservation in temperate regions: can they be predicted Agricultural <strong>and</strong><br />

<strong>Forest</strong> Entomology, 3: 233-240.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


M. Kolström et al. 2010. Is it possible to combine adaptation to climate change <strong>and</strong> maintaining of forest biodiversity<br />

690<br />

Lindner, M., 2000. Developing adaptive forest management strategies to cope with climate change.<br />

Tree Physiology, 20: 299-307.<br />

Lindner, M., 2007. How to adapt forest management in response to the challenges of climate<br />

change In: Koskela, J., Buck, A., Tessier du Cros, E. (Eds.), Climate change <strong>and</strong> forest<br />

genetic diversity: Implications for sustainable forest management in Europe. Bioversity<br />

International, Rome, Italy, pp. 31-42.<br />

Lindner, M., Garcia-Gonzalo, J., Kolström, M., Green, T., Reguera, R., Maroschek, M., Seidl, R.,<br />

Lexer, M.J., Netherer, S., Schopf, A., Kremer, A., Delzon, S., Barbati, A., Marchetti, M. <strong>and</strong><br />

Corona, P., 2008. Impacts of Climate <strong>Change</strong> on European <strong>Forest</strong>s <strong>and</strong> Options for<br />

Adaptation. Report for DG Agriculture, European <strong>Forest</strong> Institute, Joensuu, Finl<strong>and</strong>. 172 p.<br />

Maracchi, G., Sirotenko, O. <strong>and</strong> Bindi, M., 2005. Impacts of present <strong>and</strong> future climate variability<br />

on agriculture <strong>and</strong> forestry in the temperate regions: Europe. Climatic <strong>Change</strong>, 70: 117-135.<br />

McNeely, J.A., 1994. Protected areas for the 21st century: working to provide benefits to society.<br />

Biodiversity <strong>and</strong> Conservation, 3: 390-405.<br />

Moreno, J.M., 2005. Impacts on natural hazards of climatic origin, C. <strong>Forest</strong> fires risk. In: A<br />

Preliminary Assesment of the Impacts in Spain due to the Effects of Climate <strong>Change</strong>. Ministry<br />

of the Environment, Government of Spain, pp. 559-592.<br />

Rauscher, H.M., 1999. Ecosystem management decision support for federal forests in the United<br />

States: A review. <strong>Forest</strong> Ecology <strong>and</strong> Management, 114: 173-197.<br />

Salinger, M.J., Sivakumar, M.V.K. <strong>and</strong> Motha, R., 2005. Reducing Vulnerability of Agriculture <strong>and</strong><br />

<strong>Forest</strong>ry to Climate Variability <strong>and</strong> <strong>Change</strong>: Workshop Summary <strong>and</strong> Recommendations.<br />

Climatic <strong>Change</strong>, 70:341-362.<br />

Secretariat of the Convention on Biological Diversity, 2003. Interlinkages between biological<br />

diversity <strong>and</strong> climate change. Advice on the integration biodiversity consideration into the<br />

implementation of the United Nations Framework Convention on Climate <strong>Change</strong> <strong>and</strong> its<br />

Kyoto Protocol. Ad hoc Technical Expert Group on Biodiversity <strong>and</strong> Climate <strong>Change</strong>. 143 p.<br />

Spiecker, H., 2003. Silvicultural management in maintaining biodiversity <strong>and</strong> resistance of forests<br />

in Europe - Temperate zone. Journal of Environmental Management, 67: 55-65.<br />

Wermelinger, B., 2004. Ecology <strong>and</strong> management of the spruce bark beetle Ips typographus - a<br />

review of recent research. <strong>Forest</strong> Ecology Management, 202: 67-82.<br />

Woltjer, M., Rammer, W., Brauner, M., Seidl, R., Mohren, G.M.J. <strong>and</strong> Lexer, M.J., 2008. Coupling<br />

a 3D patch model <strong>and</strong> a rockfall module to assess rockfall protection in mountain forests.<br />

Journal of Environmental Management, 87: 373-388.<br />

Zebisch, M., Grothmann, T., Schröter, D., Hasse, C., Fritsch, U. <strong>and</strong> Cramer, W., 2005.<br />

Klimaw<strong>and</strong>el in Deutschl<strong>and</strong>, Vulnerabilität und Anpassungsstrategien klimasensitiver<br />

Systeme. Climate <strong>Change</strong> 08/05.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.) 2010,<br />

Instituto Politécnico de Bragança, Bragança, Portugal.


Ecosystem services from forests at the watershed scale


S. M. Carvalho-Ribeiro & T. Pinto-Correia 2010. <strong>Forest</strong>s in l<strong>and</strong>scapes: modelling forest l<strong>and</strong> cover patterns suitability<br />

692<br />

<strong>Forest</strong>s in l<strong>and</strong>scapes: modelling forest l<strong>and</strong> cover patterns suitability<br />

for meeting future dem<strong>and</strong>s for l<strong>and</strong>scape goods <strong>and</strong> services<br />

Sónia M. Carvalho-Ribeiro * & Teresa Pinto-Correia<br />

Group on Mediterranean Ecosystems <strong>and</strong> <strong>L<strong>and</strong>scapes</strong> (EPM) / Institute for<br />

Mediterranean Agrarian <strong>and</strong> Environmental Sciences (ICAAM) , University of Évora,<br />

Portugal<br />

Abstract<br />

<strong>Forest</strong>s deliver a range of l<strong>and</strong>scape functions being thus of utmost importance to study the<br />

effects of forest l<strong>and</strong> cover patterns on the provision of l<strong>and</strong>scape functions, particularly when<br />

considering both l<strong>and</strong> use <strong>and</strong> climate change scenarios. However, the explicit effects of forest<br />

spatial patterns on l<strong>and</strong>scape functions provided have not been well studied (Turner, 1989,<br />

2005). Based on the Portuguese Cávado <strong>and</strong> Sado catchments as case-studies, the overarching<br />

goal of the research project presented in this paper is to i) link forests spatial patterns with<br />

selected non-commodity functions provided by the l<strong>and</strong>scape: water flow regulation, flood<br />

prevention, carbon sequestration, recreation <strong>and</strong> territorial identity, ii) develop a new framework<br />

for prioritizing a set of l<strong>and</strong>scape functions at river basin scale, considering different scenarios,<br />

that derive from both l<strong>and</strong> use <strong>and</strong> climate change. The conceptual framework, based on the<br />

integration of different disciplines as well as the methodological construction of the work , are<br />

presented <strong>and</strong> discussed. This work is undergoing; therefore, comments <strong>and</strong> suggestions<br />

regarding methodological issues are welcomed.<br />

Keywords: watershed scale, forest l<strong>and</strong> cover, l<strong>and</strong>scape functions, goods <strong>and</strong> services<br />

provided by forests<br />

1. Introduction<br />

Goods <strong>and</strong> services provided by forests can be included in a broader category of goods <strong>and</strong><br />

services provided by the environment (de Groot et al., 2002; Slee, 2007; Turner <strong>and</strong> Daily,<br />

2008). Despite the large body of literature on ecosystem (or l<strong>and</strong>scape) functions, goods <strong>and</strong><br />

services, there is still not a clear consensus on the final definitions of these concepts (de Groot<br />

<strong>and</strong> Hein, 2007). The last report of the Millenium Ecosystem Assessement in 2005 made an<br />

attempt to bring order to the many definitions of “functions”, “goods” <strong>and</strong> “services”.<br />

Agreement was reached to define services as “the benefits people derive from ecosystems”, <strong>and</strong><br />

in order to avoid lengthy texts in scientific literature the term “services” is used, for both goods<br />

<strong>and</strong> services as well as the underlying functional processes <strong>and</strong> components of the ecosystems<br />

providing them (de Groot <strong>and</strong> Hein, 2007). Nevertheless, in other spheres, as management <strong>and</strong><br />

policy design, the term public goods is often used for identifying the non-commodity goods <strong>and</strong><br />

services that society expects from l<strong>and</strong>scapes (Cooper et al 2009). Many authors have<br />

highlighted a principal difference between the function <strong>and</strong> service. For example de Groot et al<br />

(2002:394) defined function as “...the capacity of ecosystems to provide goods <strong>and</strong> services that<br />

satisfy human needs either directly or indirectly”. Functions are seen as the actual (functional)<br />

* Corresponding author<br />

Email address: sribeiro@uevora.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. M. Carvalho-Ribeiro & T. Pinto-Correia 2010. <strong>Forest</strong>s in l<strong>and</strong>scapes: modelling forest l<strong>and</strong> cover patterns suitability<br />

693<br />

processes <strong>and</strong> components in ecosystems <strong>and</strong> l<strong>and</strong>scapes that provide or support, directly or<br />

indirectly, goods <strong>and</strong> services which benefit human welfare (de Groot <strong>and</strong> Hein, 2007).<br />

Figure 1 show that the manner in which forest services are provided (P) <strong>and</strong> benefits (B)<br />

delivered may occur in at least three different ways: 1) in situ, 2) omni-directional <strong>and</strong> 3)<br />

directional (Fisher et al., 2004). This is particular relevant for the underst<strong>and</strong>ing of spatial<br />

relations between the l<strong>and</strong>scape pattern <strong>and</strong> the functions.<br />

1. In situ. The services are provided <strong>and</strong> the benefits are realized in the same location (e.g. soil<br />

formation, provision of raw materials).<br />

2. Omni-directional. The services are provided in one location, but benefit the surrounding<br />

l<strong>and</strong>scape without directional bias (e.g. carbon sequestration, pollination)<br />

3. <strong>and</strong> 4. Directional. The service provision benefits a specific location due to the flow<br />

direction. In 3 down slope units benefit from services provided in uphill areas, for example<br />

water. In 4 the service provision unit could be coastal wetl<strong>and</strong>s (or forests) providing storm <strong>and</strong><br />

flood protection to a coastline.<br />

Figure 1. Places where services are provided (P) <strong>and</strong> benefits delivered (B).Source:<br />

Fisher et al (2004) p. 12<br />

The overarching goal of the research project to which this paper refers to, is to link l<strong>and</strong>scape<br />

pattern with l<strong>and</strong>scape function, considering particularly the role of forests for the provision of<br />

l<strong>and</strong>scape functions at the watershed scale. Preventing floods, regulating water flows,<br />

sequestering carbon <strong>and</strong> enhancing recreational activities <strong>and</strong> the role of l<strong>and</strong>scape in local<br />

identity, are likely to be important l<strong>and</strong>scape functions in the future <strong>and</strong> there is a need to model<br />

l<strong>and</strong> cover patterns able to support such l<strong>and</strong>scape functions. Because the former three functions<br />

are ecological <strong>and</strong> the latter two depend on human preferences <strong>and</strong> choices both ecological<br />

modelling <strong>and</strong> public preference methods will be used.<br />

In order to progress in the spatial analysis of l<strong>and</strong>scape functions, the Index of Function<br />

Suitability (IFS) developed by Pinto-Correia et al will be applied <strong>and</strong> tested (Pinto-Correia et<br />

al., 2009a). IFS is an innovative tool aiming at assessing the capacity of different l<strong>and</strong>scapes to<br />

provide different l<strong>and</strong>scape functions(Pinto-Correia et al., 2009a). It is based on the analysis of<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. M. Carvalho-Ribeiro & T. Pinto-Correia 2010. <strong>Forest</strong>s in l<strong>and</strong>scapes: modelling forest l<strong>and</strong> cover patterns suitability<br />

694<br />

the l<strong>and</strong> cover pattern, as expression of the l<strong>and</strong>scape composition <strong>and</strong> change. The IFS is<br />

based on the use of indicators for gauging the differences between best possible l<strong>and</strong> cover<br />

pattern for a certain function <strong>and</strong> the l<strong>and</strong> cover patterns likely to occur in contrasting scenario<br />

storylines. Therefore, it is a step further in the provision of ex-ante evaluations of the impacts of<br />

different trends of change in the provision of different functions (Fig 2). By using the same set<br />

of indicators to characterize both the best possible l<strong>and</strong> cover pattern for a certain activity <strong>and</strong><br />

the l<strong>and</strong> cover pattern likely to occur in different scenario storylines, the IFS allows for a<br />

quantification of the distance in the provision of functions, from one to another l<strong>and</strong> cover<br />

pattern. As such, it provides an indication of the suitability of l<strong>and</strong>scapes <strong>and</strong> their respective<br />

l<strong>and</strong> cover combinations, in the future, in delivering selected functions. The IFS is still under<br />

development <strong>and</strong> has so far been used exclusively for amenity functions. Throughout the<br />

present project, an improvement of the tool is expected, as well as a further adaptation to this<br />

Index, incorporating also “ecological” functions such as carbon sequestration, water flow<br />

regulation <strong>and</strong> flood protection by means of ecological modelling <strong>and</strong> expert based knowledge.<br />

Fig. 2 – The Index of Function Suitability is calculated measuring the distance between<br />

thepreferred l<strong>and</strong>scape (denominated as best possible l<strong>and</strong> cover pattern, as it depends on the<br />

existing l<strong>and</strong>s cover classes <strong>and</strong> possible combinations <strong>and</strong> compostions of those) to develop a<br />

given non-commodity function, <strong>and</strong> the l<strong>and</strong>scape being tested.<br />

2. Methodology<br />

This project uses a territorial approach looking at the Cávado <strong>and</strong> Sado catchments as Socio-<br />

Ecological Systems (Folke et al., 2005). The spatial nature of this analysis is needed, as the<br />

goods <strong>and</strong> services here considered are provided at the l<strong>and</strong>scape scale. The watershed has been<br />

defined as the unit of analysis, as it is the most comon <strong>and</strong> acknowledged level of organization<br />

of the l<strong>and</strong>scape per se, when all factors, both physical, biological <strong>and</strong> human, are considered<br />

(Selman, 2006). Both watersheds have rural-urban gradients from inl<strong>and</strong> to sea coast, <strong>and</strong> a<br />

l<strong>and</strong>scape composed of both urban, agricultural <strong>and</strong> forest l<strong>and</strong> covers, with predominance of<br />

forest <strong>and</strong> silvo-pastoral systems. Across these watersheds, the forests contribution for the<br />

lansdcape pattern <strong>and</strong> thus for human well-being certainly differ. By conducting a comparative<br />

study Sado catchment in South, Mediterrranean Portugal, <strong>and</strong> Cávado catchment in the Atlantic<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. M. Carvalho-Ribeiro & T. Pinto-Correia 2010. <strong>Forest</strong>s in l<strong>and</strong>scapes: modelling forest l<strong>and</strong> cover patterns suitability<br />

695<br />

Minho) the diversity of roles forest might play are highlighted. It is acknowledged that the<br />

l<strong>and</strong>scape functions forests contribute to will be prioritized in different ways in each of the<br />

catchments. The criteria, factors <strong>and</strong> importance of forests to different l<strong>and</strong>scape functions can<br />

be revealed by contrasting region characteristics.<br />

2.1 Identifying appropriate forest spatial patterns for different l<strong>and</strong>scape functions<br />

2.1.1. Assess the l<strong>and</strong>scape pattern, through the application of l<strong>and</strong>scape metrics for<br />

Cávado <strong>and</strong> Sado catchments. The l<strong>and</strong> cover base maps will be created in Arc/GIS <strong>and</strong><br />

following exported to a spatial pattern software in order to quantify l<strong>and</strong> cover patterns. A set of<br />

l<strong>and</strong>scape metrics (Botequilha Leitao <strong>and</strong> Ahern, 2002) will be computed in FRAGSTAT<br />

(McGarigal et al., 2002) based on the Carta de Ocupacao do Solo COS 90. By computing the<br />

l<strong>and</strong>scape metrics such as mean patch size (PS), number of patches (NP), patch density (PD)<br />

<strong>and</strong> percentage of l<strong>and</strong>scape (PLAND) for agriculture, forests (by tree species) <strong>and</strong> urban, an<br />

underst<strong>and</strong>ing of the l<strong>and</strong>scape pattern of the studied regions will be gained <strong>and</strong> a comparative<br />

analysis become possible. After quantifying the l<strong>and</strong> cover patterns within the two catchments a<br />

modelling approach will be undertaken in order to create the best l<strong>and</strong> cover pattern for different<br />

types of l<strong>and</strong>scape function. This will be carried out in two different ways as there will be focus<br />

on ecological related functions such as water flow regulation, flood protection <strong>and</strong> carbon<br />

sequestration that require expert knowledge in order to identify a best l<strong>and</strong> cover patterns within<br />

the catchments. On the contrary in the case of functions related with public preferences namely<br />

recreation <strong>and</strong> cultural identity a preferred l<strong>and</strong>scape pattern will be based on questionnaire<br />

surveys assessing preferences of different groups of users of the l<strong>and</strong>scape (hunters, farmers).<br />

2.1.2. Expert panels. The best possible l<strong>and</strong>scape patterns for the provision of l<strong>and</strong>scape<br />

functions such as water flow regulation, carbon sequestration <strong>and</strong> flood protection will be<br />

created based on expert panels at the catchment scale. “Experts” are defined as “representatives<br />

of organisations which are affected by, or which significantly affect, interactions between<br />

regional l<strong>and</strong>scapes <strong>and</strong> forests, or have skills <strong>and</strong> knowledge about the issue, or influence<br />

implementation instruments relevant to the topic”. This task will be also based on a literature<br />

review.<br />

2.1.3. Ecological modelling. Based on the data gathered suitability maps for carbon<br />

sequestration, flood protection <strong>and</strong> water regulation, will be created. L<strong>and</strong> use suitability<br />

analysis aims at identifying the most appropriate spatial pattern for l<strong>and</strong> uses according to given<br />

suitability criteria (Malczewski, 2004). Based on the ecological thresholds for different tree<br />

species suitability maps will be first created for broadleaves (oak), coniferous (pine), eucalyptus,<br />

agriculture <strong>and</strong> pastures. The l<strong>and</strong> use modelling criteria, as well as the weightings for factors<br />

will be discussed by the expert panels. These suitability maps will be combined through a<br />

Multi-Objective L<strong>and</strong> Use Allocation (MOLA) comm<strong>and</strong> in the IDRISI ANDES GIS (Eastman,<br />

2006) creating optimal l<strong>and</strong>scape patterns for carbon sequestration, water regulation <strong>and</strong> flood<br />

protection.<br />

2.1.4.) Public preferences. The best possible l<strong>and</strong> cover patterns for recreation <strong>and</strong> for cultural<br />

<strong>and</strong> identity preservation will be assessed through surveys to specific groups of l<strong>and</strong>scape users.<br />

This survey is to be done through direct enquiries based on l<strong>and</strong>scape photos used as visual<br />

stimuli for the expression of preferences. The results will be translated into the best possible<br />

l<strong>and</strong>scape pattern, adapting themethodology of Pinto-Correia et al. (2009b). Data has already<br />

been, or is being collected in research projects such as Agrorreg, Mural, Rosa (Pinto-Correia et<br />

al., 2009b) <strong>and</strong> in the work by Carvalho Ribeiro (Carvalho-Ribeiro, 2009; Carvalho-Ribeiro <strong>and</strong><br />

Lovett, 2009, 2010; Carvalho-Ribeiro et al., 2010) .<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. M. Carvalho-Ribeiro & T. Pinto-Correia 2010. <strong>Forest</strong>s in l<strong>and</strong>scapes: modelling forest l<strong>and</strong> cover patterns suitability<br />

696<br />

2.2 Developing new framework for assessing the suitability of different l<strong>and</strong>scapes in<br />

providing l<strong>and</strong>scape functions<br />

After developing best possible forest l<strong>and</strong> cover patterns for different l<strong>and</strong>scape functions a<br />

following step intends to create a framework able to prioritize a set of functions according to<br />

predefined criteria. The conceptual models is based on the premises that not all l<strong>and</strong>scapes are<br />

adequate to provide all functions being thus likely that certain areas are more suited to specific<br />

functions than areas with contrasting characteristics.<br />

The LAM L<strong>and</strong>scape Amenity Model, developed together with the IFS (Pinto-Correia 2009a)<br />

will be further developed <strong>and</strong> tested in this project.. The LAM has been designed as the<br />

modelling tool that allows for a comparison between different l<strong>and</strong> cover petterns, both those<br />

corresponding to the best possible combinations for selected functions <strong>and</strong> those resulting from<br />

identified scenarios of change, being this comparison supported on thre IFS. The l<strong>and</strong>scape<br />

pattern (created from expert knowledge models (2.1.2 <strong>and</strong> 2.1.3) <strong>and</strong> from public preferences<br />

surveys (2.1.4)) will be compared for two scenario l<strong>and</strong>scape patterns by using the LAM. In the<br />

LAM the Index of Function Suitability (IFS) evaluates the adaptatibility of a l<strong>and</strong>scape to<br />

provide a function (Pinto-Correia et al., 2009a). By using the LAM best possible l<strong>and</strong>scape<br />

patterns for different functions will be compared with scenario storylines in order to undertake<br />

an ex-ante evaluations of whether or not future forest l<strong>and</strong> cover will be able to provide a set of<br />

pre-defined l<strong>and</strong>scape functions. Also the suitability of Cávado <strong>and</strong> Sado in providing different<br />

l<strong>and</strong>scape functions will be explored <strong>and</strong> contrasted.<br />

3. Questions still to be solved<br />

When needed, <strong>and</strong> if there will be data or knowledge that allows the identification of other best<br />

possible patterns, for other functions, the evaluation path hereby presented can be exp<strong>and</strong>ed. It<br />

is not the aim here to be exhaustive the functions selected have been chosen due to their<br />

relevance in the study areas considered. In addition to the previous it is also an aim of the<br />

project to contribute for the methodological development, through a successive<br />

conceptualization, testing <strong>and</strong> improvement of research approaches able to grasp with the<br />

linkeages between l<strong>and</strong> cover <strong>and</strong> l<strong>and</strong> function independently if those functions are either<br />

ecological or based on social dem<strong>and</strong>. ,<br />

Identifying best possible l<strong>and</strong>scape patterns, or preferred patterns, can easily be criticised, as<br />

those best patterns dem<strong>and</strong>, on one side, a high degree of simplification concerning the<br />

l<strong>and</strong>scape in itself, <strong>and</strong>, on the other, they are based on the assumption that a generalised best<br />

pattern can be identified, even for functions dependent on public preferences. But still, this<br />

approach has been built up step by step taking care of the relevant literature, <strong>and</strong> has the<br />

advantage of being flexible so that it can be applied to multiple l<strong>and</strong>scapes, multiple functions,<br />

<strong>and</strong> multiple scenarios.<br />

There is still much to play regarding the complex task of linking forest spatial patterns with<br />

l<strong>and</strong>scape functions. This paper presents <strong>and</strong> explains a methodological framework able to deal<br />

with these issues. There is still uncertainty regarding the approach proposed <strong>and</strong> certainly there<br />

will be drawbacks, but the risk is taken in face of the urge imposed to the scientific community<br />

by policy makers. Policy makers are dealing with the formulation <strong>and</strong> targeting of public<br />

policies that should deal with the provision of public goods, <strong>and</strong> thus with their assessment for a<br />

given scale of analysis is of utmost importance. As previously referred this work is still in<br />

progress therefore comments, suggestions <strong>and</strong> discussion of alternative approaches will be very<br />

much appreciated.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


S. M. Carvalho-Ribeiro & T. Pinto-Correia 2010. <strong>Forest</strong>s in l<strong>and</strong>scapes: modelling forest l<strong>and</strong> cover patterns suitability<br />

697<br />

References<br />

Botequilha Leitao, A., Ahern, J., 2002, Applying l<strong>and</strong>scape ecological concepts <strong>and</strong> metrics in<br />

sustainable l<strong>and</strong>scape planning, L<strong>and</strong>scape <strong>and</strong> Urban Planning 59(2):65-93.<br />

Carvalho-Ribeiro, S. M., 2009, The role of multifunctional forests in sustainable l<strong>and</strong>scapes: A<br />

case study from Portugal, in: PhD Thesis, School of Environmental Sciences, University<br />

of East Anglia, Norwich.<br />

Carvalho-Ribeiro, S. M., Lovett, A., 2009, Associations between forest characteristics <strong>and</strong><br />

socio-economic development: A case study from Portugal, Journal of Environmental<br />

Management 90:2873-2881.<br />

Carvalho-Ribeiro, S. M., Lovett, A., 2010, Is an attractive forest well managed Correlating<br />

public preferences for forests across rural/urban gradients, Submitted to <strong>Forest</strong> Policy<br />

<strong>and</strong> Economics. Accepted for publication in April 2010.<br />

Carvalho-Ribeiro, S. M., Lovett, A., O'Riordan, T., 2010, Multifunctional forest management in<br />

Northern Portugal: Moving from scenarios to governance for sustainable development,<br />

L<strong>and</strong> Use Policy (in press) doi: 10.1016/j.l<strong>and</strong>usepol.2010.02.008<br />

de Groot, R., Hein, L., 2007, Concepts <strong>and</strong> valuation of l<strong>and</strong>scape functions at different scales,<br />

in: Multifunctional l<strong>and</strong> use: meeting future dem<strong>and</strong>s for l<strong>and</strong>scape goods <strong>and</strong> services<br />

(U. M<strong>and</strong>er, H. Wiggering, K. Helming, eds.), Springer, Berlin, pp. 14-36.<br />

de Groot, R. S., Wilson, M. A., Boumans, R. M. J., 2002, A typology for the classification,<br />

description <strong>and</strong> valuation of ecosystem functions, goods <strong>and</strong> services, Ecological<br />

Economics 41(SPECIAL ISSUE: The Dynamics <strong>and</strong> Value of Ecosystem Services:<br />

Integrating Economic <strong>and</strong> Ecological Perspectives):393–408.<br />

Eastman, J. R., 2006, IDRISI Andes Manual, Clark Labs, Clark University. Worcester. MA.<br />

USA.<br />

Fisher, B., Costanza, R., Turner, K., Morling, P., 2004, Defining <strong>and</strong> Classifying Ecosystem<br />

Services for Decision Making, in: CSERGE Working Paper EDM 07-04 (CSERGE, ed.),<br />

UEA, Norwich.<br />

Folke, C., Hahan, T., Olsson, P., Norberg, J., 2005, Adaptative governance of social- ecological<br />

systems, Annual Review Environmental Resources 30:8.1-8.33.<br />

Malczewski, J., 2004, GIS-based l<strong>and</strong>-use suitability analysis: a critical overview, Progress in<br />

Planning 62:3-65.<br />

McGarigal, K., Cushman, S. A., Neel, M. C., Ene, E., 2002, FRAGSTATS: Spatial pattern<br />

analysis program for categorical maps, Computer software program produced by the<br />

authors at the University of Massachusetts, Amherst.,<br />

www.umass.edu/l<strong>and</strong>eco/research/fragstats/fragstats.htm, 2009/03/09.<br />

Pinto-Correia, T., Machado, C., Picchi, P., 2009a, Assessing how l<strong>and</strong>scapes under different<br />

scenarios of change support amenities: the Index of Function Suitability, L<strong>and</strong> Use<br />

Policy (submitted, July 2009).<br />

Pinto-Correia, T., Machado, C., Picchi, P., Ollson, J. A., Turpin, N., Bousset, J. P., Bockstaller,<br />

C., Bezlepkina, I., 2009b, <strong>L<strong>and</strong>scapes</strong> Amenities Model (LAM), Integrated Project EU<br />

FP 6 (contract n. 010036), SEAMLESS- System for Environmental <strong>and</strong> Agricultural<br />

Modelling; Linking European Science <strong>and</strong> Society.<br />

Selman, P., 2006, Planning at the l<strong>and</strong>scape scale, Routledge, London <strong>and</strong> New York, pp. 213.<br />

Slee, B., 2007, L<strong>and</strong>scape goods <strong>and</strong> services related to forestry l<strong>and</strong> use, in: Multifunctional<br />

l<strong>and</strong> use: Meeting future dem<strong>and</strong>s for l<strong>and</strong>scape goods <strong>and</strong> services (U. M<strong>and</strong>er, H.<br />

Wiggering, K. Helming, eds.), Springer, Berlin, pp. 64-82.<br />

Turner, K. R., Daily, G. C., 2008, The Ecosystem Services Framework <strong>and</strong> Natural Capital<br />

Conservation, Environmental <strong>and</strong> Resource Economics 39:25-35.<br />

Turner, M. G., 1989, L<strong>and</strong>scape Ecology: The effect of pattern on process, Annu. Rev. Ecol.<br />

Syst. 20:171-197.<br />

Turner, M. G., 2005, L<strong>and</strong>scape Ecology: What is the state of science, Annu. Rev. Ecol. Evol.<br />

Syst. 36:319-344.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.P. Nunes et al. 2010. Impacts of wildfires on catchment hydrology<br />

698<br />

Impacts of wildfires on catchment hydrology: results from monitoring<br />

<strong>and</strong> modeling studies in northwestern Iberia<br />

João Pedro Nunes 1* , José Javier Cancelo 2 , María Ermitas Rial 1 , Maruxa Malvar 1 , Filipa<br />

Tavares 3 , Diana Vieira 1 , Frederike Schumacher 3 , Francisco Díaz-Fierros 2 , António<br />

Dinis Ferreira 4 , Celeste Coelho 1 & Jan Jacob Keizer 1<br />

1<br />

University of Aveiro, Portugal<br />

2<br />

University of Santiago de Compostela, Spain<br />

3<br />

Dresden University of Technology, Germany<br />

4<br />

Polytechnic Institute of Coimbra, Portugal<br />

Abstract<br />

Wildfires have significant impacts on vegetation cover <strong>and</strong> soil properties, which in turn can<br />

lead to large increases in runoff, erosion <strong>and</strong> nutrient export from forested hillslopes. However,<br />

the impacts of these changes at the catchment scale are still poorly understood, mostly due to<br />

the difficulty of instrumenting burnt catchments with enough speed to capture hydrological<br />

processes at the early stages of forest recovery. Hydrological modeling remains therefore an<br />

important tool to assess wildfire impacts in ungauged catchments, but most existing models<br />

cannot reproduce hydrological processes in burnt soils.<br />

This work will present results from ongoing research projects in the northwestern Iberian<br />

peninsula, including (i) multi-year runoff, erosion <strong>and</strong> soil properties monitoring in burnt <strong>and</strong><br />

unburnt hillslopes <strong>and</strong> micro-catchments (0.1 to 10 Km 2 ), providing insights into the local-scale<br />

impacts of wildfires on hydrological processes; (ii) modeling approaches to simulate<br />

hydrological processes in burnt areas at the plot <strong>and</strong> watershed scales.<br />

Keywords: hydrological modeling, burnt forests, upscaling<br />

1. Introduction<br />

<strong>Forest</strong>ed watersheds in northwestern Iberia provide important ecosystem services for water<br />

resources provisioning. According to the Millennium Ecosystem Assessment (Pereira et al.,<br />

2004), these include the regulation of runoff (preventing flash floods) <strong>and</strong> maintenance of<br />

downstream water quality. However, the planting of homogenous commercial forests has<br />

significantly increased forest fire risk in the Iberian Peninsula (Puigdefábregas, 1998); climate<br />

change is expected to increase the frequency <strong>and</strong> severity of these fires by enhancing the effect<br />

of environmental drivers behind them (e.g. Carvalho et al., 2001). This could lead to a<br />

disruption of ecosystem services during the post-fire window of disturbance (i.e. as forest<br />

vegetation regrows), leading to an increase in e.g. the likelihood of flooding or a degradation of<br />

water quality (Pereira et al., 2004). However, there is still a lack of knowledge on the effects of<br />

wildfires on hydrological processes, especially at larger spatial <strong>and</strong> temporal scales (Shakesby<br />

<strong>and</strong> Doerr, 2006).<br />

* Corresponding author.<br />

Email address: jpcn@ua.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.P. Nunes et al. 2010. Impacts of wildfires on catchment hydrology<br />

699<br />

The impact of wildfires on hydrological processes at small spatial (e.g. microplots, plots,<br />

hillslopes) <strong>and</strong> temporal scales (e.g. rainstorms, a few years) has received much attention from<br />

researchers in recent years. Shakesby <strong>and</strong> Doerr (2006) review some of the main conclusions<br />

from these studies. Wildfires reduce vegetation cover <strong>and</strong> can change soil properties, such as<br />

reducing soil aggregate stability <strong>and</strong> increasing soil water repellence. This can lead to an<br />

increase in overl<strong>and</strong> flow generation <strong>and</strong> sediment detachment, <strong>and</strong> to the formation of an<br />

extensive rill <strong>and</strong> gully system in burnt areas with additional erosion impacts. These impacts<br />

were also observed in northwestern Iberian burnt forests (e.g. Ferreira et al., 2008; Keizer et al.,<br />

2008).<br />

At larger spatial scales, especially catchments, Shakesby <strong>and</strong> Doerr (2006) report that forest<br />

fires can increase peak runoff rates, but there are few studies on consequences for total runoff.<br />

Erosion responses at this scale are more complex, as they appear to depend on smaller-scale<br />

changes to runoff <strong>and</strong> soil erosion, <strong>and</strong> are poorly studied. At larger temporal scales, studies<br />

have focused on the post-fire “window of disturbance” (c. 2 to 5 years in this region) during<br />

vegetation regrowth, but there have been few studies of long-term impacts, particularly the<br />

cumulative consequences of multiple wildfires.<br />

There is therefore a knowledge gap when upscaling measured results to larger scales <strong>and</strong><br />

impacts. Modeling has been viewed as an approach to overcome this difficulty (Shakesby <strong>and</strong><br />

Doerr, 2006). Recent comparisons of existing hydrological models with measured data in burnt<br />

areas, however, have revealed their limitations in representing the most important processes (e.g.<br />

Larsen <strong>and</strong> MacDonald, 2007) although suggesting ways to improve them. This has limited<br />

modeling applications at large spatial <strong>and</strong> temporal scales to perform watershed-scale, long-term<br />

assessments. However, this assessment is feasible provided that existing measured data can be<br />

combined with modeling tools in a coherent approach.<br />

This work will present ongoing research <strong>and</strong> results of research projects studying burnt<br />

catchment hydrology in northwestern Iberia, including:<br />

1. research at the hillslope scale conducted in central Portugal (Albergaria, Vouga basin);<br />

2. research at the micro-catchment scale in central Portugal (Mondego basin) <strong>and</strong> western<br />

Galicia (Esteiro), the latter focusing on a paired catchment study;<br />

3. modeling for meso-scale watersheds in central Portugal (Águeda basin).<br />

The results will be discussed in terms of their indication on the impacts of wildfires on<br />

watershed-scale water resources provisioning,<br />

2. Study areas<br />

The research presented in this work focuses on the northwestern Iberian Peninsula<br />

(Figure 1), a region characterized by a humid climate (aridity index, i.e. the ratio of<br />

rainfall over potential evapotranspiration, above 1) with a south-north transition from<br />

Wet Mediterranean to Maritime Temperate. There are good conditions for vegetation<br />

growth, <strong>and</strong> a large part of this region is covered by forests; in many cases, <strong>and</strong><br />

especially in Portugal, there are large areas of commercial forestry where eucalypt <strong>and</strong><br />

maritime pine are planted. However, the seasonal <strong>and</strong> interannual variability of climatic<br />

conditions, particularly the existence of a dry summer season (Figure 2), leads to the<br />

recurrence of appropriate conditions for wildfires.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.P. Nunes et al. 2010. Impacts of wildfires on catchment hydrology<br />

700<br />

3. Research at the hillslope scale<br />

This research was carried out for six hillslopes close to Albergaria, central Portugal: four burnt<br />

in 2005 (Açores 1&2, Jafafe 1&2) <strong>and</strong> two burnt in 2006 (Soutelo 1&2). All slopes were<br />

covered by eucalypt prior to the wildfires, <strong>and</strong> represent different management options<br />

including terracing <strong>and</strong> downslope ploughing. Meteorology, runoff <strong>and</strong> erosion were monitored<br />

weekly in each slope; soil surface <strong>and</strong> subsurface conditions were monitored bi-weekly using 5-<br />

point transects. Figure 3 shows some measured results for the slope Açores 1, as reported by<br />

Keizer et al. (2008). They illustrate one of the main findings of this study regarding slope-scale<br />

hydrology in burnt hillslopes. Soils in these slopes are highly water repellent immediately after<br />

fires; the repellency decreases at the start of the wet season, <strong>and</strong> reappears during the following<br />

dry season (although it can reestablish during dry spells in the wet season, as shown in the<br />

figure for the winter of 2007). Repellency <strong>and</strong> soil water storage are closely related, <strong>and</strong><br />

reasonable amounts of rainfall falling on repellent soils might not lead to an increase of soil<br />

water. Consequently, rainfall events at the start of the wet season (or during other periods with<br />

high soil water repellency) might lead to increased runoff rates when compared with events in<br />

the middle <strong>and</strong> end of this season; this has the potential for increasing flood risks at the<br />

catchment scale which might be difficult to predict if repellency is not taken into account.<br />

4. Watershed-scale research: micro- <strong>and</strong> meso-scale catchments<br />

Research at the watershed scale is shown for two micro-catchment sites, Colmeal (central<br />

Portugal) <strong>and</strong> Esteiro (Galicia); <strong>and</strong> one meso-catchment site, Águeda (central Portugal).<br />

Colmeal is a micro-catchment (60 ha) which burnt almost completely in 2008; it was<br />

instrumented with a number of slope-scale plots, while runoff <strong>and</strong> sediment data was also<br />

collected at the catchment outlet in a multi-scale sampling approach. Preliminary results<br />

indicate that runoff <strong>and</strong> erosion are concentrated in the first months after the wildfire, at the start<br />

of the vegetation disturbance <strong>and</strong> when soil water repellency is high. They also indicate a nonlinear<br />

relationship between slope-scale <strong>and</strong> catchment-scale runoff, as the latter also depends on<br />

subsurface flow <strong>and</strong> saturation-excess runoff generation close to the water line.<br />

The Esteiro fire occurred in 2007; two paired micro-catchments were instrumented, each with<br />

around 450 ha (Figure 5). The Maior catchment was partially burnt while the Arestiño<br />

catchment was not. Preliminary results indicate that runoff was more irregular in the burnt<br />

catchment, with higher values in the winter immediately following the wildfires <strong>and</strong> lower<br />

values in the subsequent summer (close to zero). These results might indicate an impact of<br />

wildfires on increasing the irregularity of water resources provisioning, although differences in<br />

the underlying geology of both catchments might also play a role in differentiating the<br />

hydrological regimes. The Águeda fire occurred in 1986, <strong>and</strong> burned c. 25% of the upper<br />

Águeda river basin. Given the scarcity of hydrological data for this wildfire (runoff was only<br />

measured at the outlet), a modeling approach is being used to estimate the impacts on runoff <strong>and</strong><br />

water balance on subcatchments with different burnt areas. The approach relies on the SWAT<br />

model (Neitsch et al., 2005), capable of simulating a number of processes inside meso-scale<br />

catchments, including water balance, surface <strong>and</strong> subsurface runoff, soil erosion, nutrient<br />

exports <strong>and</strong> vegetation regrowth (including human management operations). The current<br />

modeling approach relies on simulating 5-10 years in both burnt <strong>and</strong> unburnt conditions,<br />

therefore evaluating the different impacts for dry <strong>and</strong> wet years as well as assess the relative<br />

importance of wildfires <strong>and</strong> climate variability for watershed-scale water yield.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.P. Nunes et al. 2010. Impacts of wildfires on catchment hydrology<br />

701<br />

References<br />

Carvalho, A.C., Carvalho, A., Mir<strong>and</strong>a, A.I., Borrego, C., <strong>and</strong> Rocha, A., 2001. Climatic<br />

<strong>Change</strong> <strong>and</strong> the Fire Weather Risk. In: M. Brunet India <strong>and</strong> D. López Bonillo (Eds.).<br />

Detecting <strong>and</strong> Modelling Regional Climate <strong>Change</strong>. Berlin, Germany: Springer-Verlag:<br />

555-565.<br />

Ferreira, A.J.D., Coelho, C.O.A., Ritsema, C.J., Boulet, A.K., <strong>and</strong> Keizer, J.J., 2008. Soil <strong>and</strong><br />

water degradation processes in burned areas: Lessons learned from a nested approach.<br />

Catena 74: 273-285.<br />

Keizer, J.J., Doerr, S.H., Malvar, M.C., Prats, S.A., Ferreira, R.S.V., Oñate, M.G., Coelho,<br />

C.O.A., <strong>and</strong> Ferreira, A.J.D., 2008. Temporal variation in topsoil water repellency in two<br />

recently burnt eucalypt st<strong>and</strong>s in north-central Portugal. Catena 74: 192–204.<br />

Larsen, I.J., <strong>and</strong> MacDonald, L.H., 2007. Predicting postfire sediment yields at the hillslope<br />

scale: testing RUSLE <strong>and</strong> Disturbed WEPP. Water Resources Research 43: W11412.<br />

Neitsch, S.L., Arnold, J.G., Kiniry, J.R., <strong>and</strong> Williams, J.R., 2005. Soil <strong>and</strong> Water Assessment<br />

Tool Theoretical Documentation (version 2005). Agricultural Research Service, United<br />

States Department of Agriculture, Temple, Texas, 494 p.<br />

Pereira, H.M, Domingos, T., <strong>and</strong> Vicente, L. (editors), 2004. Portugal Millennium Ecosystem<br />

Assessment: State of the Assessment Report. Centro de Biologia Ambiental, Faculdade de<br />

Ciências da Universidade de Lisboa, Lisbon, Portugal, 68 p.<br />

Puigdefábregas, J., 1998. Ecological impacts of global change on dryl<strong>and</strong>s <strong>and</strong> their<br />

implications for desertification. L<strong>and</strong> Degrad. Develop., 9: 393-406.<br />

Shakesby, R.A., <strong>and</strong> Doerr, S.H., 2006. Wildfire as a hydrological <strong>and</strong> geomorphological agent.<br />

Earth-Science Reviews 74: 269-307.<br />

Figure 1: location of the study areas in the Iberian Peninsula, over a map of the aridity index (left).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.P. Nunes et al. 2010. Impacts of wildfires on catchment hydrology<br />

702<br />

Anadia, Portugal (1961-1990)<br />

Santiago de Compostela, Spain (1971-2000)<br />

Rainfall (mm)<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Temperature (ºC)<br />

Rainfall (mm)<br />

300<br />

250<br />

200<br />

150<br />

PP<br />

100<br />

Tavr<br />

50<br />

0<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Temperature (ºC)<br />

PP<br />

Tavr<br />

Jan<br />

Feb<br />

Mar<br />

Apr<br />

May<br />

Jun<br />

Jun<br />

Jul<br />

Jul<br />

Aug<br />

Sep<br />

Oct<br />

Nov<br />

Dec<br />

Jan<br />

Feb<br />

Mar<br />

Apr<br />

May<br />

Jun<br />

Jul<br />

Aug<br />

Sep<br />

Oct<br />

Nov<br />

Dec<br />

Figure 2: monthly climate normal for weather stations in central Portugal (left) <strong>and</strong> Galicia (right).<br />

Rainfall/PET (mm)<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Rainfall<br />

PET<br />

Soil water repellency<br />

(severity)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

26<br />

10<br />

24 7<br />

215<br />

199<br />

23 6<br />

20<br />

133<br />

18 8<br />

15<br />

29<br />

12<br />

19<br />

10<br />

24<br />

20<br />

28 9<br />

30<br />

278<br />

225<br />

27<br />

12<br />

26<br />

23 7<br />

213<br />

18<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Soil water storage (mm)<br />

SWR Q1 to Q3<br />

SWR Median<br />

Soil w ater<br />

9 10 11 12 1 2 3 4 5 6 7 9 10 11 1 2 3 4 5 6<br />

2005 2006 2007<br />

Figure 3: measured rainfall <strong>and</strong> potential evapotranspiration (top), <strong>and</strong> soil water storage <strong>and</strong> repellency<br />

(bottom) for the Açores 1 slope; soil water repellency is indicated as the severity class associated with the<br />

result of the Molarity of Ethanol Droplet (MED) test, <strong>and</strong> the shaded area represents the interquartile area<br />

(between the 1 st <strong>and</strong> 3 rd quartiles).<br />

Figure 4: Colmeal micro-catchment <strong>and</strong> wildfire limits.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.P. Nunes et al. 2010. Impacts of wildfires on catchment hydrology<br />

703<br />

Figure 5: Esteiro study area, including the Maior (burnt) <strong>and</strong> Arestiño (unburnt) micro-catchments.<br />

Figure 6: Águeda meso-scale catchment, showing the burnt area in the 1986 wildfires.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Management <strong>and</strong> conservation of Mediterranean forest<br />

l<strong>and</strong>scapes


P.M. Fern<strong>and</strong>es et al. 2010. Testing the fire paradox: is fire incidence in Portugal affected by fuel age<br />

705<br />

Testing the fire paradox: is fire incidence in Portugal affected by fuel<br />

age<br />

Paulo M. Fern<strong>and</strong>es 1,2* , Carlos Loureiro 1,2 , Marco Magalhães 2 & Pedro Ferreira 2<br />

1<br />

Centro de Investigação e de Tecnologias Agro-Ambientais e Biológicas (CITAB),<br />

UTAD, Apartado 1013, 5001-801 Vila Real, Portugal<br />

2<br />

Departamento de Ciências Florestais e Arquitectura Paisagista, UTAD, Apartado 1013,<br />

5001-801 Vila Real, Portugal<br />

Abstract<br />

A well-known fire paradox states that fire exclusion <strong>and</strong> the resulting fuel accumulation leads to<br />

larger fires. To test this assumption for Portugal we analyzed fire frequency through survival<br />

analysis for the period of 1975-2008. The median fire-free interval is relatively short (12-16<br />

years) but the hazard of burning increases exponentially with time since fire, denoting a fuel-age<br />

dependent system. Time-dependency did not change with fire size class, implying that young<br />

fuels delay l<strong>and</strong>scape fire spread even under extreme meteorological conditions. Fire size <strong>and</strong><br />

maximum fire size tend respectively to be less variable <strong>and</strong> lower where fire recurrence is<br />

higher. Hence, fuels exert a short-term but effective control on l<strong>and</strong>scape fire spread. Our<br />

findings show that crown fire regimes can be time-dependent <strong>and</strong> support fuel treatments as key<br />

component of fire management.<br />

Keywords: burn probability, fire regime, shrubl<strong>and</strong>, Mediterranean-type ecosystems<br />

1. Introduction<br />

Controversy concerning the environmental drivers of wildfire incidence has gained momentum<br />

in recent years. If fire occurrence is time-dependent then exogenous factors such as weather will<br />

play a relatively minor role on fire incidence, which to a great extent will be controlled by the<br />

existing mosaic of fuel ages (Minnich 1983; Minnich <strong>and</strong> Chou 1997). This can be expressed as<br />

a “fire paradox” where fire suppression efforts will lead to increasingly larger fires as fuel<br />

builds up in the l<strong>and</strong>scape. However, the opposing view is now prevailing in the literature <strong>and</strong><br />

defends that fuel age is unrelated to the likelihood of wildfire in crown-fire ecosystems such as<br />

conifer forests <strong>and</strong> Mediterranean-type shrubl<strong>and</strong>s, particularly under extremely dry <strong>and</strong> windy<br />

weather conditions (Keeley et al. 1999; Moritz et al. 2004; Keeley <strong>and</strong> Zedler 2009). Because<br />

the question is difficult to be addressed empirically, more <strong>and</strong> more studies resort to simulation<br />

modeling (e.g. Finney et al. 2007; Cary et al. 2009). Fire frequency analysis has been proposed<br />

to judge whether vegetation is fire-adapted or not (Polakow et al. 1999) <strong>and</strong> has been used to<br />

better underst<strong>and</strong> <strong>and</strong> describe how the fire return interval <strong>and</strong> fuel age affect fire incidence<br />

(Moritz 2003; Moritz et al. 2004; O’Donnell et al. 2008; Van Wilgen et al. 2010). Obviously,<br />

this debate has profound political implications, namely on the recognition (or not) of fuel<br />

management as a valuable component of fire management.<br />

While extreme fire seasons significantly affected the country in 2003 <strong>and</strong> 2005, extensive tracts<br />

of Portugal are under a relatively frequent fire regime, especially in mountains <strong>and</strong> plateaus<br />

* Corresponding author. Tel.: +351 259350885 - Fax: +351 259350480<br />

Email address: pfern@utad.pt<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.M. Fern<strong>and</strong>es et al. 2010. Testing the fire paradox: is fire incidence in Portugal affected by fuel age<br />

706<br />

dominated by shrubl<strong>and</strong> where fire is a tool routinely used in traditional l<strong>and</strong> management. The<br />

existing spatial <strong>and</strong> temporal information can thus be analyzed to detect relationships between<br />

fuel age <strong>and</strong> fire incidence. Our objective is to analyze contemporary fire history data to<br />

determine how fire recurrence, hence fuel age, relates with burn probability in Portugal.<br />

2. Methodology<br />

The study area is the entire Portuguese mainl<strong>and</strong>, 89 x 10 3 km 2 . The analysis of burn probability<br />

was based on all wildfire events with an area equal to or larger than 10 ha occurring in Portugal<br />

from 1998 to 2008, i.e. during a 11-year period. We used the mapped fire history of the<br />

Portuguese <strong>Forest</strong> Service <strong>and</strong> GIS software to process the spatial information. Fire recurrence<br />

maps were created for each year under analysis, fire recurrence being the number of fires<br />

experienced by each 625-m 2 pixel (25 x 25 m) since 1975. For the area within each individual<br />

fire perimeter we have determined an area-weighed mean fire recurrence <strong>and</strong> the corresponding<br />

fire return interval (FRI). Then we used survival analysis by fitting a two-parameter Weibull<br />

function to the FRI distribution based on maximum likelihood. We have modelled FRI from<br />

individual fire events, rather than from patches defined by their unique fire history, because the<br />

former offers the possibility of assessing if the time-dependency of burn probability changes<br />

with fire size, i.e. with weather conditions.<br />

A fire interval distribution can be described in a cumulative form, F(t), the probability of fire<br />

occurrence before or at time t, <strong>and</strong> as a probability density, f(t), which reflects the frequency of<br />

burning in a given time interval (e.g. Moritz 2003; Moritz et al. 2009):<br />

F(t) = 1-exp[-(t/b) c ] (1)<br />

f(t) = (ct c-1 /b c ) exp[-(t/b) c ] (2)<br />

where t > 0, b > 0 <strong>and</strong> c ≥ 0. Parameters b <strong>and</strong> c have ecological meaning. The scale parameter<br />

(b) has the dimensions of time <strong>and</strong> is the typical fire return interval (FRI) that will be surpassed<br />

36.79% of the time. The shape parameter (c) is dimensionless <strong>and</strong> describes the change in fire<br />

probability through time. The negative exponential distribution is a special case of the Weibull<br />

model that corresponds to c = 1. The probability of burning increases with time when c > 1,<br />

increasing linearly when c = 2 <strong>and</strong> exponentially when c > 2. Vegetation types that are<br />

simultaneously fire-prone <strong>and</strong> fire-dependent are expected to have high c values (Polakow et al.<br />

1999) <strong>and</strong> c > 4 qualifies a fire regime as highly age-dependent (Moritz 2003). The Weibull<br />

median fire-free interval (MEI) gives a probabilistic estimate of fire-return intervals for<br />

asymmetrical fire interval distributions (Grissino-Mayer 1999). The “hazard of<br />

burning“ function λ(t) gives the instantaneous probability of a fire occurring in a specific time<br />

interval <strong>and</strong> is useful to measure how time since the last fire event affects the subsequent<br />

likelihood of burning:<br />

λ(t) = ct c-1 /b c (3)<br />

Data was truncated to consider only those fires whose majority of pixels had burnt at least twice<br />

since 1975, thus reducing the existing asymmetry that otherwise would have biased parameters<br />

b <strong>and</strong> c (Moritz 2003); note that fire incidence in Portugal was quite low before the 1970s. The<br />

Weibull model was fitted separately to two sub-divisions of Portugal − northern <strong>and</strong> centraleastern<br />

(N-CE) <strong>and</strong> central-western <strong>and</strong> southern (CW-S) − mainly on the basis of their relative<br />

fire incidence, high in N-CE <strong>and</strong> relatively low in CW-S (Verde <strong>and</strong> Zêzere 2010).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.M. Fern<strong>and</strong>es et al. 2010. Testing the fire paradox: is fire incidence in Portugal affected by fuel age<br />

707<br />

3. Results<br />

The Portuguese fire atlas contains 10197 burned patches with ≥10 ha in the 1998-2008 period,<br />

corresponding to a sum of 1.65 x 10 6 ha (of which 58% have burned just once) <strong>and</strong> a mean<br />

annual burned surface of 1.5 x 10 5 ha. The largest patch has 66071 ha <strong>and</strong> the mean <strong>and</strong> median<br />

patch sizes were 162 <strong>and</strong> 31 ha, respectively. The fire frequency analysis was restricted to the<br />

2950 fire scars whose majority of pixels had burned at least twice before.<br />

Examination of FRI data (Figure 1a) indicates a sigmoidal (S-shaped) trend in the cumulative<br />

fire probability curves, which indicates a fuel age effect on burn probability. Resistance to fire is<br />

nevertheless of relatively short duration. Region N-CE displays a steep slope (starting at age 6)<br />

with the result that more than 75% of the total burning is reached by age 15. Region CW-S<br />

exhibits low burn likelihood for about 11 years, followed by a sharp increase up to age 20.<br />

Table 1 <strong>and</strong> Figure 1b respectively present the fitted Weibull model parameters for N-CE <strong>and</strong><br />

CW-S <strong>and</strong> their corresponding hazard functions. The typical fire interval length is higher in<br />

CW-S. The likelihood of burning grows exponentially with time in both sub-divisions of the<br />

country, CW-S having a especially high degree of age dependency (c = 4). In N-CE the hazard<br />

of burning (% per year) increases by a factor of three from age 4 (1.9%) to age 7 (5.7%), <strong>and</strong><br />

again at age 12 (16.8%), while in CW-S those rates are reached approximately at ages 7, 11 <strong>and</strong><br />

15, respectively; N-CE <strong>and</strong> CW-S hazards of burning are equal at age 23. The likelihood of<br />

burning is below that of an age-independent system (i.e. c = 1) for 10 years in N-CE <strong>and</strong> 18<br />

years in CW-S.<br />

Table 1 also includes a fire frequency analysis as a function of wildfire size class (< 100 ha,<br />

100-499 ha, 500-999 ha, ≥ 1000 ha) for N-CE. The Weibull model b <strong>and</strong> c coefficients were<br />

quite similar between fire size classes, with overlapping 95% confidence intervals.<br />

Consequently, extreme fire weather does not change the fuel-age dependency of burn<br />

probability in Portugal. Note also that fuel age dependency is higher in the drier CW-S subdivision.<br />

Table 2 displays additional information regarding the size of the burnt patches for the entire data<br />

set (n=10197). Size variability <strong>and</strong> maximum size are higher for burnt patches belonging to<br />

larger size classes. Again, this points to a fuel age buffering effect against extreme fire weather.<br />

Figure 1: Empirical probability of fire occurrence (a) <strong>and</strong> fitted Weibull hazard of burning (b) for the N-<br />

CE <strong>and</strong> CW-S sub-divisions of Portugal.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.M. Fern<strong>and</strong>es et al. 2010. Testing the fire paradox: is fire incidence in Portugal affected by fuel age<br />

708<br />

4. Discussion<br />

Considering the fire return intervals involved, this study reflects primarily the fire regime of<br />

shrubl<strong>and</strong> <strong>and</strong> regenerating forest. The results are coherent with the fuel dynamics described for<br />

shrubl<strong>and</strong> in northern (Fern<strong>and</strong>es <strong>and</strong> Rego 1998) <strong>and</strong> central Portugal (Fern<strong>and</strong>es et al. 2000)<br />

which indicate increased flammability with time since fire; depending on shrubl<strong>and</strong> type, fine<br />

fuel loads in these systems are at near-equilibrium 9 to 26 years after fire.<br />

While the typical fire return interval is shorter in N-CW, age-dependency is more apparent in<br />

CE-S Portugal. The differences can be explained by distinct fire environments in terms of<br />

physiography, vegetation <strong>and</strong> l<strong>and</strong> use, as well as by differences in climate <strong>and</strong> ignition density.<br />

Topography is rougher <strong>and</strong> vegetation types − namely atlantic <strong>and</strong> sub-atlantic shrubl<strong>and</strong> types<br />

(Ulex, Erica, Pterospartum, Cytisus) <strong>and</strong> Pinus pinaster <strong>and</strong> Eucalyptus globulus forest − are<br />

more flammable in N-CE than in CW-S Portugal. Also, because of drier climate, growth rates<br />

<strong>and</strong> fuel accumulation are expected to be slower over much of CW-S in comparison with N-CE.<br />

Mediterranean shrubl<strong>and</strong> types prevail in CW-S (limestone maquis <strong>and</strong> garrigue, Cistus), <strong>and</strong><br />

because their canopy is poor in dead fuels, exhibit fire behaviour patterns in relation to weather<br />

that are distinct from atlantic <strong>and</strong> sub-atlantic shrubl<strong>and</strong>s <strong>and</strong> that may increase the agedependency<br />

of fire hazard. Density of fire ignitions is also substantially higher in N-CE, where<br />

traditional fire use at short-return intervals for pastoral purposes is common in the mountains.<br />

Higher fire recurrence may promote a fuel complex with a significant grass component that will<br />

burn more readily, hence facilitating more frequent fire.<br />

Table 1: Weibull model parameters (<strong>and</strong> 95% confidence intervals) <strong>and</strong> median fire-free interval (MEI)<br />

for the fire frequency analysis. Fire size class data respect to the N-CE sub-division.<br />

Data set n MEI b c<br />

N-CE 2862 12.1 13.7 (13.5-13.9) 3.0 (2.9-3.1)<br />

CW-S 88 15.5 17.0 (16.1-17.9) 4.1 (3.5-4.7)<br />

Fire size class<br />

10-99 ha 2131 12.3 13.9 (13.7-14.1) 3.0 (2.9-3.1)<br />

100-499 ha 594 11.7 13.2 (12.8-13.6) 3.0 (2.8-3.1)<br />

500-999 ha 88 11.2 12.6 (11.7-13.5) 3.1 (2.6-3.6)<br />

≥ 1000 ha 49 12.2 13.7 (12.4-15.1) 3.2 (2.6-3.8)<br />

Table 2: Burnt patch size statistics by fuel age class.<br />

Fuel age class (yrs.) n Median (ha) CV (%) Max. (ha)<br />

3-4 26 50.2 117.7 379<br />

5-6 197 43.4 146.1 1100<br />

7-8 418 48.3 190.4 2196<br />

9-12 1043 44.0 222.9 5487<br />

13-19 1681 41.8 351.1 13119<br />

≥20 6832 27.3 876.6 66071<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.M. Fern<strong>and</strong>es et al. 2010. Testing the fire paradox: is fire incidence in Portugal affected by fuel age<br />

709<br />

In comparison with other Mediterranean regions, the fire-free interval is shorter in Portugal than<br />

in southern California chaparral (Moritz 2003; Moritz et al. 2004) <strong>and</strong> in southwestern Australia<br />

shrubl<strong>and</strong> (O’Donnell et al. 2008), <strong>and</strong> is comparable with South Africa fynbos (Van Wilgen et<br />

al. 2010). In contrast to our findings, shrubl<strong>and</strong> systems in other Mediterranean regions are<br />

generally characterized by weak (Moritz 2003; Moritz et al. 2004; Van Wilgen et al. 2010) to<br />

moderate (O’Donnell et al. 2008) age-dependency. While several factors other than vegetation<br />

type <strong>and</strong> fuel hazard can affect fuel age dependency − fire suppression, l<strong>and</strong>form, l<strong>and</strong>scape<br />

fragmentation, frequency <strong>and</strong> severity of extreme weather events – <strong>and</strong> be involved in the<br />

results, this study shows that not all crown fire regimes are independent of fuel age. The usual<br />

expectation is that extreme weather conditions will prevail over, or will cancel the effect of fuel<br />

on l<strong>and</strong>scape fire spread (Fern<strong>and</strong>es <strong>and</strong> Botelho 2003; Moritz 2003; Keeley <strong>and</strong> Zedler 2009),<br />

but such pattern did not emerge in the analysis. The results imply that the fuel age effect on burn<br />

probability is weather-independent, which is highly relevant for fire management, as it increases<br />

the expectations of effective fuel treatment performance under unfavourable weather scenarios.<br />

Consequently, this study supports a policy where l<strong>and</strong>scape-level, strategically-placed<br />

prescribed burning treatments are an integral component of fire management.<br />

Acknowledgments<br />

This study has been funded by the EC project FIRE PARADOX (FP6-018505).<br />

References<br />

Cary, G.J., Flannigan, M.D., Keane, R.E., Bradstock, R.A., Davies, I.D., Lenihan, J.M., Li, C.,<br />

Logan, K.A. <strong>and</strong> Parsons, R.A., 2009. Relative importance of fuel management, ignition<br />

management <strong>and</strong> weather for area burned: evidence from five l<strong>and</strong>scape-fire-succession<br />

models. International Journal of Wildl<strong>and</strong> Fire, 18: 147–156.<br />

Fern<strong>and</strong>es, P.M. <strong>and</strong> Botelho, H.S., 2003. A review of prescribed burning effectiveness in fire<br />

hazard reduction. International Journal of Wildl<strong>and</strong> Fire, 12: 117-128.<br />

Fern<strong>and</strong>es, P. <strong>and</strong> Rego, F.C., 1998. <strong>Change</strong>s in fuel structure <strong>and</strong> fire behaviour with heathl<strong>and</strong><br />

aging in Northern Portugal. In: Proceedings of the 13th Conference on Fire <strong>and</strong> <strong>Forest</strong><br />

Meteorology. International Association of Wildl<strong>and</strong> Fire: 433-436.<br />

Fern<strong>and</strong>es, P., Ruivo, L., Gonçalves, P., Rego, F. <strong>and</strong> Silveira, S., 2000. Dinâmica da<br />

combustibilidade nas comunidades vegetais da Reserva Natural da Serra da Malcata. In:<br />

Livro de Actas do Congresso Ibérico de Fogos Florestais. <strong>Escola</strong> <strong>Superior</strong> Agrária de<br />

Castelo Branco: 177-186.<br />

Finney, M.A., Seli, R.C., McHugh, C.W., Ager, A.A., Bahro, B. <strong>and</strong> Agee, J.K., 2007.<br />

Simulation of long-term l<strong>and</strong>scape-level fuel treatment effects on large wildfires.<br />

International Journal of Wildl<strong>and</strong> Fire, 16: 712-727.<br />

Grissino-Mayer, H.D., 1999. Modeling fire interval data from the American Southwest with the<br />

Weibull distribution. International Journal of Wildl<strong>and</strong> Fire, 9: 37-50.<br />

Keeley, J.E., <strong>and</strong> Zedler, P.H., 2009. Large, high-intensity fire events in southern California<br />

shrubl<strong>and</strong>s: debunking the fine-grain age patch model. Ecological Applications, 19: 69-<br />

94.<br />

Keeley, J.E., Fotheringham, C.J. <strong>and</strong> Morais, M., 1999. Reexamining fire suppression impacts<br />

on brushl<strong>and</strong> fire regimes. Science, 284: 1829-1832.<br />

Minnich, R.A., 1983. Fire mosaics in Southern California <strong>and</strong> Northern Baja California.<br />

Science, 219: 1287-1294.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


P.M. Fern<strong>and</strong>es et al. 2010. Testing the fire paradox: is fire incidence in Portugal affected by fuel age<br />

710<br />

Minnich, R.A. <strong>and</strong> Chou, Y.H., 1997. Wildl<strong>and</strong> fire patch dynamics in the chaparral of southern<br />

California <strong>and</strong> northern Baja California. International Journal of Wildl<strong>and</strong> Fire, 7: 221-<br />

248.<br />

Moritz, M.A., 2003. Spatiotemporal analysis of controls on shrubl<strong>and</strong> fire regimes: age<br />

dependency <strong>and</strong> fire hazard. Ecology, 84: 351-361.<br />

Moritz, M.A., Keeley, J.E., Johnson, E.A. <strong>and</strong> Schaffner, A.A., 2004. Testing a basic<br />

assumption of shrubl<strong>and</strong> fire management: how important is fuel age Frontiers in<br />

Ecology <strong>and</strong> the Environment, 2: 67-72.<br />

Moritz, M.A., Tadashi, J.M., Miles, L.J., Smith, M.M. <strong>and</strong> de Valpine, P., 2009. The fire<br />

frequency analysis branch of the pyrostatistics tree: sampling decisions <strong>and</strong> censoring in<br />

fire interval data. Environmental <strong>and</strong> Ecological Statistics, 16: 271-289.<br />

O'Donnell, A., McCaw, L., Boer, M. <strong>and</strong> Grierson, P., 2008. Wildfire return intervals in semiarid<br />

Southern Western Australia: effects of fuel age <strong>and</strong> spatial structure. In: Fire,<br />

Environment <strong>and</strong> Society: from Research into Practice: the International Bushfire<br />

Research Conference Incorporating the 15th AFAC Conference. Melbourne, Bushfire<br />

CRC: 520.<br />

Polakow, D., Bond, W., Lindenberg, N. <strong>and</strong> Dunne, T.T., 1999. Ecosystem engineering as a<br />

consequence of natural selection: methods for testing Mutch's hypothesis from a<br />

comparative study of fire hazard rates. In: Proceedings of Australian Bushfire<br />

Conference: 321-328.<br />

Verde, J.C. <strong>and</strong> Zêzere, J.L., 2010. Assessment <strong>and</strong> validation of wildfire susceptibility <strong>and</strong><br />

hazard in Portugal. Natural Hazards Earth Systems Science, 10: 485-497.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


Interfaces <strong>and</strong> interactions between forest <strong>and</strong> agriculture in<br />

rural l<strong>and</strong>scapes


E. Andrieu et al. 2010. When forests are managed by farmers<br />

712<br />

When forests are managed by farmers:<br />

Implications of farm practices on forest management<br />

E. Andrieu 1,2 , A. Sourdril 3 , G. du Bus de Warnaffe 4 , M. Deconchat 1,2 , G. Balent 1,2<br />

1 INRA; UMR 1201 DYNAFOR, F-31326 Castanet Tolosan, France<br />

2 Université de Toulouse, UMR DYNAFOR, INRA / INP-ENSAT, BP 32607, 31326<br />

Castanet Tolosan, France<br />

3 INRA, MONA, 65 Boulevard de Br<strong>and</strong>ebourg, 94205 Ivry-sur-Seine Cedex, France<br />

4 Arbre et Bois Conseil, Domaine du Chalet, route de Chalabre, 11 300 Limoux, France<br />

Abstract<br />

Farm forests, i.e. forests managed by farmers, are important components of French l<strong>and</strong>scapes.<br />

Farmers, who do not have knowledge in sylviculture in general, harvest them for firewood <strong>and</strong><br />

timberwood, but also for hunting, mushroom harvesting or grazing. The social <strong>and</strong> ecological<br />

functions of these woods call for a better underst<strong>and</strong>ing of their management. These private<br />

woods are mainly small (< 25 ha) <strong>and</strong> thus are not submitted to French management regulations.<br />

We present the conclusions of three multidisciplinary long term studies, in south west of France,<br />

based on historical, social <strong>and</strong> technical analyses of the particularities of these woodlots. Results<br />

showed that the traditional social system (“house-centered system”) is still influencing forest<br />

management, despite its loosing of importance. Woodlots are parts of the agricultural systems<br />

but some cultural features limit the implementation modern forestry practices. The roles of farm<br />

forests have to be considered on a larger l<strong>and</strong>scape scale perspective.<br />

Keywords: Rural forest, private forest, management pratices, SIG, ethnology<br />

1. Introduction<br />

In France, 28 % of the territory is forested <strong>and</strong> around 75% of this forested area (i.e. 10 millions<br />

of ha) belongs to a multitude of private owners (around 3.5 millions). A large part of these<br />

private woods are farm forests that are forested areas managed <strong>and</strong> used by farmers, whatever<br />

are the legal system <strong>and</strong> the structure of the st<strong>and</strong>s (Norm<strong>and</strong>in 1996). They represent 17% of<br />

French private forest <strong>and</strong> up to 50% in several departements in South-West (Cinotti <strong>and</strong><br />

Norm<strong>and</strong>in 2002). Despite their important cultural <strong>and</strong> ecological functions in agricultural<br />

l<strong>and</strong>scapes (Balent et al. 1996, Sourdril 2008) <strong>and</strong> their potential role in the sustainable<br />

development of rural areas, management modalities of farm forests are still poorly known.<br />

Moreover, their surface area is decreasing because of their conversion into common private<br />

forest, due to sales <strong>and</strong> inheritance to non-farmers (Cinotti <strong>and</strong> Norm<strong>and</strong>in 2002): farming <strong>and</strong><br />

forestry are thus more <strong>and</strong> more disconnected (Larrère <strong>and</strong> Nougarède 1990; Cardon 1999).<br />

A characteristic of French private forest is the small size of the properties: half of them<br />

have an area of less than 25 ha. Yet regularly obligations do not impose particular forest<br />

management in these small properties, unlike in larger ones for which a management schedule<br />

(Plan Simple de Gestion, regulatory document that is both a guide for forest management <strong>and</strong> a<br />

traceability document) approved by the Centre Régional de la Propriété <strong>Forest</strong>ière (public<br />

institution supporting private sylviculturists to manage their forests) is compulsory. As a<br />

consequence, management in small private woods is not described in any document or<br />

inventory that hampers the study of practices <strong>and</strong> the build up of forest management history.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Andrieu et al. 2010. When forests are managed by farmers<br />

713<br />

The objective of this paper is to present an original combination of retrospective<br />

mapping from aerial photographs of farm forest management for 60 years with anthropological<br />

analysis of the drivers of forest management <strong>and</strong> forestry practices. We thus have to analyze<br />

forestry practices thanks to aerial photographs <strong>and</strong> to interviews analysis. Interviews were also<br />

used to assess forest representational systems that are linked to practices (Lemonnier 1994), <strong>and</strong><br />

the relationships among stakeholders concerned by farming <strong>and</strong> forestry (social networks).<br />

2. Methodology<br />

2.1 Study sites<br />

Studied woods are located in the Long Term Ecological <strong>and</strong> Sociological Research platform «<br />

Vallées et Coteaux de Gascogne » in southwestern France, at ~ 60 km south-west of Toulouse,<br />

in 5 rural parish territories (43° 16’ N, 48° 43' E, 200 - 400 m a.s.l.). This hilly region is<br />

characterized by a temperate climate with oceanic <strong>and</strong> slight Mediterranean influences. Two<br />

types of soil occur in the study site: superficial calcareous soils (superficial terrefort) <strong>and</strong> noncalcareous<br />

acid molasse (brown acid <strong>and</strong> brown washed soils) (CRAMP 1995). The region is<br />

not densely populated <strong>and</strong> is still largely agricultural. The dominant tree species of the woods<br />

are sessile oak (Quercus petraea Lieblein) <strong>and</strong> pedunculate oak (Q. robur L.), mixed with<br />

hornbeam (Carpinus betulus L.) (Cabanettes <strong>and</strong> Guyon 1994). Wood management system is<br />

traditionally coppice (with or without st<strong>and</strong>ard trees) each 30 years by clear cutting mostly to<br />

produce fire wood <strong>and</strong> some st<strong>and</strong>ard trees can be cut to be sold or used as timber wood.<br />

2.2 History of forest logging<br />

We selected a set of woods (total surface area ~ 100 Ha) with a common history, i.e. they were<br />

all included in a whole larger wood two centuries ago that has been fragmented. Aerial<br />

photographs are the most appropriate data that allow reconstructing logging history at a fine<br />

scale. We chose 7 missions conducted by the French National Geographical Institute <strong>and</strong> by the<br />

French National <strong>Forest</strong> Inventory (1942, 1953, 1962, 1977, 1984, 1996, 2006) in order to obtain<br />

a regular temporal sequence. The photographs were analyzed using optical stereoscopy in order<br />

to detect <strong>and</strong> date cuttings. We circumscribed polygons defined by their cover classes, based on<br />

tree height (Bakis <strong>and</strong> Bonin 2000, Muraz et al. 1999): cuttings (C), young coppice (R) <strong>and</strong><br />

mature st<strong>and</strong>s (M) (see De Warnaffe et al. 2006). These cover classes were digitalized on aerial<br />

photographs georeferenced with ArcGis®. We adapted the most frequently used photointerpretation<br />

method (regressive photo-interpretation method, which consists in digitizing from<br />

the most recent to the older photograph (Muraz et al. 1999): indeed, at the fine scale of our<br />

study, superimposed digitalized maps do not fit perfectly because of inherent imprecision of<br />

georeferencing process, which induces artefactual edge changes (see Andrieu et al. 2008).<br />

Resulting maps were crossed in a SIG to obtain a synthesis map summarizing logging history.<br />

Between 1942 <strong>and</strong> 2006, we assessed temporal changes (Mann-Kendall test for<br />

monotonic trends) in cutting system : cutting number, cutting total surface area per year, cutting<br />

mean surface area, ratio of areas coppiced with vs. without st<strong>and</strong>ards, cutting shapes complexity<br />

(Area Weighted Mean Shape Index, <strong>and</strong> Area Weighted Mean Patch Fractal Dimension).<br />

Matches between the shapes of cuttings <strong>and</strong> cadastral parcels were estimated visually. To test<br />

whether both ecological factors <strong>and</strong> forest accessibility can influence the number of cutting<br />

between 1942 <strong>and</strong> 2006 (elevation, slope <strong>and</strong> aspect, distance for streams, forest edge <strong>and</strong> ways,<br />

geology <strong>and</strong> soil type), we built log-linear regression model (log link function) on a r<strong>and</strong>om<br />

sample of 2000 points generated in the study area. Significance of effects was assessed by<br />

comparing deviance reduction of nested models with χ² tests (anova function, stats package, R)<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Andrieu et al. 2010. When forests are managed by farmers<br />

714<br />

<strong>and</strong> contrasts tests were used to compare regression coefficient estimates among levels. We<br />

analyzed wood maturation through changes in the proportion of surface area of cuttings, mature<br />

<strong>and</strong> immature st<strong>and</strong>s. To assess spatio-temporal stability of mature st<strong>and</strong> habitat, we built cores<br />

of mature st<strong>and</strong> areas with potential edge effect of 10 m to 50 m for each period: maps were<br />

pooled in order to detect areas that were continuously mature since 1942.<br />

2.3 Analysis of forest management practices<br />

In order to analyze forest management practices we selected 7 woods with different surface<br />

areas (0.7 – 11.2 ha) on which we had rebuild the history of logging. They belong to 11 private<br />

owners: 4 active farmers, 7 retired farmers <strong>and</strong> 7 non-farmers. We used semi-directive<br />

interviews to analyze practices, know-how <strong>and</strong> ethnobotanical knowledge of private owners,<br />

<strong>and</strong> to define social networks. Both interviews <strong>and</strong> visits in the woods allowed rebuilding<br />

« mental-maps » of cuttings realized by the owners or their family. At home, we asked the<br />

owner to draw on an empty map the shape of the cutting procedure as far as he could remember.<br />

In the wood, the owner completed the information obtained (seeing logging signs helped the<br />

owner remind) <strong>and</strong> modify the previous map drawn. These “as told” practices (mental maps)<br />

where crossed with “observed” practices (from aerial pictures) in a GIS. During the interviews,<br />

the owner was also asked to describe the nature of the cutting procedure, cuttings people<br />

involved, season <strong>and</strong> equipment used, st<strong>and</strong>ards maintained, <strong>and</strong> what use was made of the<br />

wood that was cut since 1938. From these descriptions, two hypotheses have been tested. First,<br />

retired farmers are traditionally the managers of the woodlots after the transfer of the farm<br />

ownership <strong>and</strong> management to their son (Nougarède 1999). We tested thus the hypothesis of<br />

such a separation between knowledge <strong>and</strong> / or practices of forestry <strong>and</strong> farming which can be<br />

led by this transfer (Nougarède 1999; Cardon 1999). Second, the inheritance of woodlots to<br />

non-farmers could lead to a separation of forestry practices between farmers <strong>and</strong> non-farmers,<br />

each having their own knowledge <strong>and</strong> social networks. We tested thus whether the management<br />

of woodlots by non-farmers is disconnected or not from agriculture.<br />

3. Results <strong>and</strong> Discussion<br />

Whereas they constitute a large part of forested area in France, modalities <strong>and</strong> history of<br />

management of small private woods remained widely misunderstood because of the difficulty to<br />

collect <strong>and</strong> analyze historical information. Based on the analysis of historical documents <strong>and</strong> of<br />

semi-directive interviews, our studies show clearly how complex in space <strong>and</strong> time forest<br />

management by private owners can be.<br />

3.1 Comparing mental maps <strong>and</strong> photo-interpreted aerial photography: benefits from the<br />

comparison between “as told” <strong>and</strong> “observed” practices<br />

First, there was a good agreement between the “as told” <strong>and</strong> “observed” practices. Three types<br />

of disagreement were detected: (1) cutting operation is reported at the time of the interview<br />

without being identified on the aerial photograph, generally due to an high density of the<br />

st<strong>and</strong>ards; (2) cutting operation is detected by the aerial photograph but not reported at the time<br />

of the interview, due to an incomplete memory; <strong>and</strong> (3) rarely a great difference in the cutting<br />

areas for a given date or cutting date very different for a given area which can point out the<br />

difficulty of the informant to legend the “mental map”. Since it is not possible to verify the<br />

memory of the owners, we should place greater faith in the aerial photographs to detect the<br />

cutting places; however cutting places choices <strong>and</strong> cutting procedures cannot be understood<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Andrieu et al. 2010. When forests are managed by farmers<br />

715<br />

without the memory of the owners. The combination of the two methods is therefore required to<br />

underst<strong>and</strong> forest management through space <strong>and</strong> time.<br />

Second, photo-interpretation showed some evolutions of management practices between<br />

1942 <strong>and</strong> 2006. In our study site, the management consisted in numerous small cuttings (< 1<br />

ha), mainly coppiced with st<strong>and</strong>ard (65% of the cuttings). The type of management (coppice<br />

with or without st<strong>and</strong>ards), the shape of the cuts, <strong>and</strong> their spatial localization (i.e. their<br />

aggregation) remained stable through time. However, even if the number of cutting remained<br />

stable, their mean size decrease since around 1980 (0.43-0.63 ha - depending on the year -<br />

before 1980, 0.16-0.20 ha after), which caused a decrease of total cut surface area (3.8 to 1.4<br />

ha). This pattern associated to general information collected on rural life changes in the<br />

interviews indicate that the origin of the decrease of total cut surface area is not rural<br />

depopulation <strong>and</strong> the subsequent ab<strong>and</strong>onment of forested parcels. It is very likely to be either<br />

the energetic transition to fossil fuel that decreased the needs in firewood (which is the main use<br />

of wood in our study site), or the lack of time available for forest management activities (<strong>and</strong><br />

more widely the management of semi-natural habitats) in a context of change of farmer<br />

activities during the last decades. An important ecological consequence of the reduction of<br />

wood harvest is the maturation of the st<strong>and</strong>s. In our study site, st<strong>and</strong> age structure changed<br />

drastically <strong>and</strong> we are now assisting to a homogenization of st<strong>and</strong> maturity toward mature trees:<br />

in 1942, 6 % of wood surface area was mature <strong>and</strong> 88% immature whereas in 2006 59 % was<br />

mature <strong>and</strong> 39% immature. Such changes of age structure can lead to a change of biodiversity<br />

patterns, like the replacement of assemblages dominated by early successional species by<br />

assemblages dominated by late successional species.<br />

Photo-interpretation showed that the main part of the woods has been submitted to a<br />

moderate harvest intensity <strong>and</strong> has been cut once (27 % of surface area), twice (38 %) or three<br />

times (16 %) (Fig. 1). In appearance, around 15% of the wood surface area has never been cut<br />

since 1942. However these areas tally mainly with ancient ab<strong>and</strong>oned fields or parts which have<br />

been already cut just before 1942. When these areas are excluded, only 3% of wood surface<br />

areas that were mature in 1942 have never been cut between 1942 <strong>and</strong> 2006. These preserved<br />

areas are scarce <strong>and</strong> small, particularly when a small edge effect of 10 m is attributed to them (9<br />

isolated patches, surface area


E. Andrieu et al. 2010. When forests are managed by farmers<br />

716<br />

l<strong>and</strong>); (2) some stakeholders can manage forests without being owner or decision-maker. We<br />

thus have to consider together property, decision <strong>and</strong> action: our typology is then owner / nonowner,<br />

decision-maker / non-decision-maker <strong>and</strong> action performer / action non-performer.<br />

These characteristics are non exclusive, for example a performer can occasionally make<br />

decisions or an owner decision-maker can occasionally carry out some forest activities. The<br />

term of « manager » is then meaningless in our study context. Therefore, even if the owner is<br />

not a farmer - anymore or at all - farmers can be involved in the forest practices at some point.<br />

Following the idea that forest can be hold by retired father, while the son deal with the<br />

farm, we try to underst<strong>and</strong> with our new typology the different roles shared by fathers <strong>and</strong> sons<br />

in forest management. Four father / son couples have been interviewed. All fathers were retired<br />

farmer, owners <strong>and</strong> decision makers, two were performers (the two other were occasional<br />

performers). No son (active farmers) was owner but all were decision-makers together with<br />

their fathers, three were the main performers <strong>and</strong> one was occasional performer. Retired farmer<br />

always helps their son in the farm. When the woods belong to the father, a strong relationship<br />

between agricultural <strong>and</strong> forestry practices remains, due to the implication of both father <strong>and</strong><br />

son on both agricultural <strong>and</strong> forest l<strong>and</strong>s. Even if they sometimes disagree about the way forests<br />

have to be managed, the management of the forests belonging to retired farmers is not<br />

dissociated from agricultural practices. Moreover most retirees keep considering <strong>and</strong> are<br />

considered to be as farmer as active farmers themselves. In the process of transmission of the<br />

property, if the forest can be separated from the farm for a while because being kept by the<br />

father while the farm is given to the son, this process is never definitive <strong>and</strong> the forest will<br />

automatically be transmitted to the son at the death of his father; it is finally never detached<br />

from the farm except on the paper.<br />

What happens when non farmers are owners of forest l<strong>and</strong> Five non-farmers were<br />

owners <strong>and</strong> two were non-owner but were performers or decision-makers. Whereas non-farmers<br />

have no professional link with farming activities, we found that their forestry practices were<br />

strongly influenced by farming activities because (A) they can have relatives who are farmers<br />

<strong>and</strong> they integrate their practices, know-how <strong>and</strong> representations of the forest; (B) forests are<br />

embedded in an agricultural matrix so logging can be dependent of agricultural practices (for<br />

example, a farmer can ask his neighbors to manage the forest edge adjacent to his field); (C)<br />

techniques can percolate through social networks (neighbors, friends).<br />

To conclude, these multidisciplinary studies illustrate that the specificity of rural forest<br />

management, particularly the high spatio-temporal variability of cuttings, the strong links<br />

between forestry <strong>and</strong> farm practices even when owners are “non-farmers”, <strong>and</strong> the organization<br />

of activities within farming families.<br />

References<br />

Andrieu, E., du Bus de Warnaffe, G., Ladet, S., Heintz, W., Sourdril, A., Deconchat, M., 2008.<br />

Mapping the felling history in small-sized woodl<strong>and</strong>s, Revue <strong>Forest</strong>ière Française, 60:<br />

667-676.<br />

Bakis, H., Bonin, N., 2000. La Photographie aérienne et spatiale. Que sais-je, PUF, Paris.<br />

Balent, G., 1996. La forêt paysanne dans l’espace rural. Biodiversité, paysages, produits.<br />

Etudes et Recherche sur les Systèmes Agraires et le Développement, INRA Editions, 29.<br />

Cabanettes, A., Guyon, J.P., 1994. Relations entre gestion et structure dans les systèmes boisés<br />

d’exploitations agricoles. In Colloque « Agriculteurs, agricultures et forêts », Cemagref<br />

et INRA, Paris, 12-13 décembre 1994.<br />

Cardon, P., 1999. Un capital dormant. La transmission patrimoniale de la forêt paysanne en<br />

Franche-Comté. Terrain, 32.<br />

Cinotti, B., Norm<strong>and</strong>in, D., 2002. Exploitants agricoles et propriété forestière : où est passée la<br />

« forêt paysanne » Revue <strong>Forest</strong>ière Française, 4: 311-327.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


E. Andrieu et al. 2010. When forests are managed by farmers<br />

717<br />

Chambre Régionale d’Agriculture de Midi-Pyrénées, 1995. Les gr<strong>and</strong>es ensembles morphopédologiques<br />

de la région Midi-Pyrénées.<br />

De Warnaffe, G., Deconchat, M., Ladet, S., Balent, G., 2006. Variability of cutting regimes in<br />

small private woodlots of south-western France. Annals of <strong>Forest</strong> Science, 63: 915-927.<br />

Larrère, R., Nougarède, O., 1990. La forêt dans l’histoire des systèmes agraires. De la<br />

dissociation à la réinsertion. Cahiers d’Economie et de sociologie rurales, 15-16: 11-38.<br />

Lemonnier, P., 1994. Choix techniques et représentations de l’enfermement chez les Anga de<br />

Nouvelle-Guinée. In Latour B et Lemonnier P (éd.), De la préhistoire aux missiles<br />

balistiques, La découverte, Paris, pp. 253-272.<br />

Muraz, J., Durrieu, S., Labbe, S., Andreassian, V., Tangara, M., 1999. Comment valoriser les<br />

photos aériennes dans les SIG. Ingénieries, 20: 39−57.<br />

Nougarède, O., 1999. La transmission de la forêt paysanne. Les bois dans la vie de la famille<br />

agricole. In: Bois et forêts des agriculteurs, Actes du colloque, Clermont-Ferr<strong>and</strong>, 20-21<br />

oct. 1999, Cemagref Editions, pp. 309-333.<br />

Sourdril, A., 2008. Territoire et hiérarchie dans une société à maison Bas-Commingeoise :<br />

Permanence et changement. Des bois, des champs, des prés (Haute-Garonne), PhD<br />

dissertation Univ. Paris X.<br />

Figure 1: Synthesis map of cutting number between 1942 <strong>and</strong> 2006 (dark green = 0, light green = 1,<br />

yellow = 2, orange = 3, red = 4 <strong>and</strong> more, black lines = delimitation of different cutting events).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. Geni et al. 2010. A framework for characterizing convergence <strong>and</strong> discrepancy in rural forest management<br />

718<br />

A framework for characterizing convergence <strong>and</strong> discrepancy in rural<br />

forest management in tropical <strong>and</strong> temperate environments<br />

D. Genin 1* , Y. Aumeerudy-Thomas 2 , G. Balent 3,4 , G. Michon 5<br />

1 IRD, Laboratoire Population, Environnement Développement, UMR151, Marseille,<br />

France<br />

2 CNRS, Centre d’Ecologie Fonctionnelle et Evolutive, Montpellier, France<br />

3 INRA, UMR DYNAFOR, Toulouse, France<br />

4 Université de Toulouse, UMR DYNAFOR, Toulouse<br />

5 IRD, UR199, Montpellier, France<br />

Abstract<br />

Rural forests are forests that are more or less formally appropriated, managed, shaped or rebuilt<br />

by rural communities, who have developed refined local knowledge <strong>and</strong> practices related to<br />

their use <strong>and</strong> perpetuation. Based on detailed monographs, we compared eleven situations of<br />

rural forests both from developing <strong>and</strong> developed countries, localized within a high diversity of<br />

ecological environments (humid tropics, dry forests, temperate forests) as well as regarding<br />

socio-economical <strong>and</strong> public policies characteristics. Data were pooled within a common<br />

analysis chart <strong>and</strong> processed by means of multivariate analyses. Results show that some<br />

variables are characteristic of all rural forests, such as multiple-use, tree species diversity,<br />

ecosystem stability, or patrimonial functions. Other results point out some specificities of<br />

particular rural forests, depending on the main use of single out of several tree species,<br />

importance of NTFPs, l<strong>and</strong> ownership <strong>and</strong> management, <strong>and</strong> the magnitude of public action.<br />

This framework aims at better characterizing these particular forests in order to think about<br />

alternative forest management policies.<br />

Keywords: rural forests, local practices, local knowledge, multivariate analyses,<br />

1. Introduction<br />

<strong>Forest</strong>s in various parts of the world are not only places for wood production or environmental<br />

services (Myers, 1988; Ros-Tonen et al., 2005). They also play several roles for local people<br />

<strong>and</strong> are deeply integrated within diversified livelihood systems (Pretzsch, 2005; Agrawal, 2007).<br />

Rural forests are all forests more or less formally managed, shaped, transformed or rebuilt by<br />

rural communities, integrated within farming systems which structure l<strong>and</strong>scapes <strong>and</strong> rural<br />

territories. These forests have not only contributed to sustain local livelihoods but are an entire<br />

part of local patrimony <strong>and</strong> identity. They are indeed based on trans generational knowledge <strong>and</strong><br />

know-how that have contributed to social reproduction. However, particularities of these rural<br />

forests are not well defined since they encompass highly diversified situations <strong>and</strong> have not<br />

been of interest for classical forest science. Very few authors have explored the intrinsic<br />

characteristics of rural forests. Balent (1995) talked about peasant forest with examples from the<br />

French situation. Michon et al. (2007) coined a new paradigm for integrating local<br />

communities’ forestry into tropical forest science with the notion of domestic forests. These<br />

authors argued <strong>and</strong> illustrated that domestic forest is “a forest for living, a forest that integrates<br />

* Corresponding author.<br />

Email address: Genin@univ-provence.fr<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. Geni et al. 2010. A framework for characterizing convergence <strong>and</strong> discrepancy in rural forest management<br />

719<br />

production <strong>and</strong> conservation with social, political, <strong>and</strong> spiritual dimensions” at local level, <strong>and</strong><br />

constitutes an entity which has to be dissociated from the more classical forests.<br />

There is a real need for a better characterization of these rural forests. Which indicators can be<br />

addressed to underst<strong>and</strong> the specific relationships that have evolved between rural people <strong>and</strong><br />

the forests they have shaped Are there general features found both in northern <strong>and</strong> southern<br />

countries As complex socio-ecological systems, how do social <strong>and</strong> ecological factors<br />

jointly determine their nature Can the concept of rural forest be an empirical model to put<br />

sustainable development in practice<br />

The research program POPULAR “Public policies <strong>and</strong> farmers’ local management of trees <strong>and</strong><br />

forests: sustainable alliance or dupery” (www2.toulouse.inra.fr/popular/), supported by the<br />

French National Research Agency, aims at exploring the links between public policies <strong>and</strong> local<br />

uses <strong>and</strong> management of small-scale forests by rural people, at sharing experiences both from<br />

North <strong>and</strong> South, <strong>and</strong> at going further in the characterization of this almost orphan research<br />

object that constitutes rural forest. As part of this project, we report here on a comparative study<br />

that attempted to show the general characteristics <strong>and</strong> particularities of rural forests from<br />

examples taken both from North (five sites in France) <strong>and</strong> South (Cameroon, India, Indonesia<br />

<strong>and</strong> two sites in Morocco).<br />

2. Material <strong>and</strong> Methods<br />

The eleven sites were chosen in order to cover a large array of situations both ecologically<br />

(humid tropical forest, dry forests, temperate forests) <strong>and</strong> concerning human population pressure,<br />

socio-economic conditions <strong>and</strong> public policies related to forest management.<br />

In France, different cases were developed within a context of centralized forest management <strong>and</strong><br />

both agricultural modernization <strong>and</strong> decline : 1) common small-scale forests privately managed<br />

by traditional farmers from south-west part of France (Gascogne) (Deconchat et al., 2007): 2<br />

<strong>and</strong> 3) the domestic chestnut forests in Corsica <strong>and</strong> Cevennes (South of France) which<br />

experienced several phases of ab<strong>and</strong>onment <strong>and</strong> renovation <strong>and</strong> where confrontations between<br />

local dynamics <strong>and</strong> sustainable development policies are recurrent (Michon & Sorba, 2010); 4)<br />

An example of a multifunctional management of a tree out of forest in the Pyrenees (southwestern<br />

France) which coevolved with changing agricultural systems <strong>and</strong> l<strong>and</strong> use<br />

management : The ash tree (Mottet et al., 2007); <strong>and</strong> 5) the Languedoc Garrigues, a case of<br />

agriculture ab<strong>and</strong>onment where local initiatives emerge to control parts of the territory, such as<br />

truffle forestry (Therville, 2009).<br />

In Morocco, two situations representative of semi arid environment were analyzed : a secular<br />

endogenous community-based forest management named agdal in the High Atlas mountains<br />

(Cordier & Genin, 2008); the argan forest, located in the South West of the Country, with a<br />

unique a semi-arid climate (annual rainfall around 350 mm). Argan tree (Argania spinosa) is an<br />

endemic tree which provides highly valued oil for alimentary <strong>and</strong> cosmetic uses, <strong>and</strong> where<br />

there is large historic trajectory of forest management both from local populations <strong>and</strong> forestry<br />

services (Simenel et al. 2009).<br />

In Southern Cameroon, even if forest ownership is public, management of forest is usually<br />

performed by local communities who partly depend upon the extraction of forest resources for<br />

their livelihood (Lescuyer, 2007).<br />

In the Western Ghâts of India, different situations coexist such as commercial agroforests,<br />

sacred forests both managed by local people, <strong>and</strong> reserved forests managed by forestry services<br />

but where collective l<strong>and</strong> use rights still continue. Two situations were analyzed in this paper:<br />

agroforests <strong>and</strong> reserved forests (Hinnewinkel et al., 2008).<br />

In Indonesia, there is a large experience on indigenous communities’ utilisation of forest<br />

resources. Since the 1970s, the relevance of local management systems for forest science,<br />

conservation <strong>and</strong> development have become well-recognized facts.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. Geni et al. 2010. A framework for characterizing convergence <strong>and</strong> discrepancy in rural forest management<br />

720<br />

In each forest case, an interdisciplinary attempt was made in order to provide a detailed<br />

description of the rural forest, taking into account stakeholders, resources, practices involved,<br />

<strong>and</strong> ecological aspects <strong>and</strong> dynamics. A common analysis grid was previously built in order to<br />

produce a corpus of comparable data. The general aim was to characterize the socio-ecological<br />

systems by putting emphasis on interactions between ecological structures <strong>and</strong> functioning, use<br />

rules <strong>and</strong> intensity of uses of natural resources both in time <strong>and</strong> space. We encompassed five<br />

main themes: 1) physical <strong>and</strong> ecological characterization; 2) Actors <strong>and</strong> use rules; 3) Uses <strong>and</strong><br />

functions of forests <strong>and</strong> forest resources; 4) Naturalist, technical, organisational, spiritual <strong>and</strong><br />

political knowledge linked with the uses of trees <strong>and</strong> forests; 5) Main dynamics <strong>and</strong> challenges<br />

related to forested areas.<br />

In a first step, monographs were analyzed <strong>and</strong> similarities or differences were pointed out by a<br />

qualitative analysis of the results within themes above mentioned. On the basis of this material<br />

<strong>and</strong> team researchers’ expertise, a comparative data base was built, composed of 58 variables.<br />

We rated the outcomes for each variable on a five-point scale: inexistent or very low (1), low or<br />

poorly important (2), neutral or average (3), high or important (4), very high or very important<br />

(5). Assessments were done by a reduced researchers’ group, <strong>and</strong> codification result discussed<br />

with the all team members of the POPULAR project. This process was effective for stimulating<br />

discussions, facilitating consistent <strong>and</strong> comparable information, <strong>and</strong> scoring of the variables. All<br />

variables were treated as having uniform weight, <strong>and</strong> were analysed together through Multiple<br />

Correspondence Analysis (MCA, see Benzecri, 1973). MCA was performed in two stages: first<br />

by fitting a cloud of points in a multidimensional vector subspace, <strong>and</strong> second, by setting a<br />

metric structure on this space. This analysis provided a non-parametric description of the<br />

relationships between modalities of variables <strong>and</strong> an indication of their importance rather than a<br />

measure of significance. It allowed the possibility of treating together both qualitative <strong>and</strong><br />

quantitative data. We performed a Hierarchical Cluster Analysis of both variables <strong>and</strong> case<br />

studies using their scores along the main MCA axes, Euclidian distance <strong>and</strong> Ward linkage<br />

function. The Pearson correlation test was used in order to evaluate similarities between sites,<br />

<strong>and</strong> between variables. Data treatments were carried out using the STATBOX software.<br />

3. Results <strong>and</strong> discussion<br />

The main difficulty in studying rural forests is to deal with the complexity <strong>and</strong> interrelationships<br />

of uses, practices, functions <strong>and</strong> representations associated with forest resources <strong>and</strong> wooded<br />

l<strong>and</strong>scapes. In this sense, a global descriptive approach of this complexity can constitute a first<br />

step in order to filter the main aspects governing the functioning of these systems. The set of<br />

case studies we undertook here, though far from being exhaustive, gives an interesting overview<br />

of this diversity in terms of ecosystem types <strong>and</strong> forest resources, forms of organization <strong>and</strong><br />

operation of local communities, role of public authorities in forest management, <strong>and</strong> functions<br />

devoted to forest areas.<br />

MCA reduction of the initial table in a few number of dimensions led to a synthetic<br />

representation of both forest case <strong>and</strong> influential variables that made possible a comparison of<br />

the similarities <strong>and</strong> dissimilarities between the forest cases. The first four MCA axes covered<br />

more than 50% of the overall observed variance, which is a relatively high score for a<br />

qualitative data set.<br />

The first axis made it possible to discriminate the tropical forests (Cameroon, Indian reserved<br />

forests <strong>and</strong> to a lesser extent Indonesia significantly correlated to little to average domestication<br />

of trees, very strong rate of tree cover associated with a very strong tree species diversity, uses<br />

related to timbering <strong>and</strong> the non timber forest products (NTFP), <strong>and</strong> a strong control of the rules<br />

at the collective level) from the Corsican <strong>and</strong> Cevennes chestnut groves <strong>and</strong> truffle growing<br />

significantly correlated to highly transformed ecosystem, little fragmentation of the l<strong>and</strong>scape,<br />

private l<strong>and</strong>ownership associated with little influence of collective institutions, strong use for<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. Geni et al. 2010. A framework for characterizing convergence <strong>and</strong> discrepancy in rural forest management<br />

721<br />

firewood, very strong technical know-how, <strong>and</strong> control of the trees <strong>and</strong> strong transformations<br />

of the social systems .<br />

The second axis allows discriminating the small private forest of Gascogne, the ash tree case<br />

<strong>and</strong> the Indian Reserved forest from the Argan forest <strong>and</strong> Indonesian agroforests. The first<br />

group of forests is correlated to private l<strong>and</strong>ownership except for the Indian reserved forest,<br />

strong compliance with the collective rules (often because they do not exist!), poor commercial<br />

economic functions associated with poor economic issues, little economic valorisation of the<br />

products extracted from the forest, <strong>and</strong> a weak forest/agriculture integration. The second one to<br />

strong domestication of the trees, little fragmentation of the l<strong>and</strong>scape, mixed l<strong>and</strong>ownership<br />

status associating private, collective <strong>and</strong> public ones, strong set of rules established at collective<br />

level <strong>and</strong> associated with a rather strong control of the rules by the State, <strong>and</strong> common use of<br />

non woody forest products.<br />

The third axis allows to discriminate truffles <strong>and</strong> small private forests of Gascogne,<br />

characterized by a domestication at the tree st<strong>and</strong> rather than at the individual tree level, few<br />

new actors intervening who claim or not of being partisans of sustainable development, little<br />

transformation of the forest products, a weak home consumption of forest products <strong>and</strong> a very<br />

little transformation of the social systems, from Corsican <strong>and</strong> Cevennes chestnut cases<br />

characterized by the importance of the use of the forest for animal husb<strong>and</strong>ry, a strong dynamic<br />

of the practices associated with strong transformation of the social systems, a high<br />

transformation of the products resulting from the forest <strong>and</strong> the importance of new actors<br />

intervening in the use of the forest.<br />

On the fourth axis, cases with a high significance were found the Agdals of the High Atlas <strong>and</strong><br />

Indian agro-forests. For the first case, outst<strong>and</strong>ing variable modalities were those related to the<br />

collective l<strong>and</strong>ownership status, the influence of the collective rules for access, uses <strong>and</strong><br />

management of forests; modalities such as the strong use of foliage as fodder, the function of<br />

the rural forest as reserve/safety in cases of climatic hazards, a high home consumption<br />

associated with a very weak commercial activity with the forest products were also significant.<br />

For the Indian agro forests, associated modalities of variables were a very strong<br />

forest/agriculture integration, associated with an agricultural use of the forest <strong>and</strong> the use of<br />

NTFP, a l<strong>and</strong>ownership mainly private, but a very strong political know-how suggesting a<br />

structured local organization; also an increase of surperficie.<br />

The Hierarchical Cluster analysis on the forest cases provided a classification of the studied<br />

rural forests in five groups: High Atlas (group 1); Cameroun, Indonesia, Indian Reserved<br />

<strong>Forest</strong>s (group 2); Indian agro forests (group 3); Corsica, Cevennes, Argan forest (group 4)<br />

Gascogne, truffles, ash tree (group 5).<br />

It pointed out similarities found in tropical forests, globally (group 2), the ones found where a<br />

single tree species represents the main characteristic <strong>and</strong> uses of the forest (chestnut groves,<br />

argan forest) (group 4), <strong>and</strong> the ones with a traditional forest <strong>and</strong> tree management found in<br />

France mainly characterized by almost completely private decisions <strong>and</strong> practices, with very<br />

weak external interventions. Two situations remained individualized: Agdals in the High Atlas<br />

where strong traditional collective rules <strong>and</strong> uses conform the overall forest management, <strong>and</strong><br />

Indian agro forests where there is a mix between individual, collective <strong>and</strong> state interventions,<br />

<strong>and</strong> a high diversity within the transect between natural to highly transformed forest.<br />

The Hierarchical Cluster analysis on the variables data set provided a partition within seven<br />

classes:<br />

Class 4 represents the overall common characteristics of the rural forests (multi-use,<br />

home consumption, importance of use of wood (firewood <strong>and</strong> timber), tree species diversity,<br />

stability of ecosystem, important patrimonial functions as well as their stakes in the construction<br />

of the territories). All the studied forests present important scores for these variables (> 3.5 on<br />

average).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. Geni et al. 2010. A framework for characterizing convergence <strong>and</strong> discrepancy in rural forest management<br />

722<br />

Classes 1 <strong>and</strong> 3 are characteristics of group 4 (Corsica, the Cevennes, Argan forest).<br />

They correspond to rural forests based on the use of a single tree species (chestnut or Argan) for<br />

which the knowledge related to control <strong>and</strong> the domestication of the individuals are important,<br />

as well as the practices of transformation <strong>and</strong> product valorization.<br />

Class 2 represents the reactive variables with the group 5 (Gascogne, truffle, ash tree)<br />

which represents the French small-scale forests integrated to traditional farming systems, very<br />

weakly influenced by the forest public policies <strong>and</strong> managed at the family level to meet various<br />

needs <strong>and</strong> amenities (timber <strong>and</strong> firewood, mushrooms, hunting places, etc).<br />

Class 5 gathers the major variables characterizing tropical rural forests (groups 2 <strong>and</strong> 3),<br />

it said the importance of timbering, an important biodiversity, the use of various non woody<br />

forest products (resins, fruits, medicinal plants, etc).<br />

Class 7 reflects the variables for which the State is very present, at the same time in the<br />

management of the forests <strong>and</strong> its relations with the local populations (High Atlas <strong>and</strong> India<br />

reserved forest).<br />

Class 6 gathers the significant variables linked with the traditional rural forests found in<br />

developing Countries (groups 1, 2 <strong>and</strong> 3) where the influence of diversified collective<br />

institutions (being customary or more or less legally formalized) is critical in the management<br />

of forest areas <strong>and</strong> where these areas have important functions reserve/safety for the livelihood<br />

systems.<br />

Conclusion<br />

In this first attempt to put evidence of what could be the specificities of rural forests, we can<br />

advance that they are, in essence, multi functional areas, of long term usefulness both for<br />

extracting goods, accommodating livelihoods to local environment, taking control of l<strong>and</strong>scapes<br />

<strong>and</strong> territories, <strong>and</strong> building a representation’s world intimately linked to local culture <strong>and</strong><br />

knowledge. Diversity logically characterizes rural forests which are an on-going result of this<br />

multifaceted shaping; it has then to be undertaken in the light of the functioning of rural<br />

societies. As Michon et al. (2007) argued, rural forests present two universal features that could<br />

characterise them. The first one concerns the local managers who are usually farmers.<br />

Management practices are closely related to agriculture (sensus lato) <strong>and</strong> range from local<br />

interventions in the forest ecosystem, to more intensive types of forest cultivation, <strong>and</strong><br />

ultimately to permanent forest plantation showing high similitude with classical agricultural<br />

practices. The second one concerns the continuum of planted forests with the natural forest, in<br />

matters of vegetation’s structure <strong>and</strong> composition, as well as economical traits <strong>and</strong> ecosystems<br />

services. Arnold (1977) claimed that uses-oriented forest management by local people <strong>and</strong> tree<br />

planting can be explained as being one or more of four categories of response to dynamics of<br />

rural change: to maintain supplies of tree products as production from off-farm tree stocks<br />

decline; to meet growing dem<strong>and</strong> for tree products; to help maintain agricultural productivity in<br />

face of declining soil quality; <strong>and</strong> to contribute to risk management <strong>and</strong> securization of the<br />

overall functioning of farming systems. We would like to add: an early perception by local<br />

people of specific forest resources’ potentialities to enhance their livelihoods, the dem<strong>and</strong> for a<br />

local l<strong>and</strong> control both for extraction <strong>and</strong> long term management, <strong>and</strong> a fundamental structural<br />

element of the way local societies perceive their surrounding world <strong>and</strong> their humanity. Hence,<br />

rural forests constitute also a critical element of the biocultural sphere of local societies which<br />

determine in a significant part, both structure of ecosystems <strong>and</strong> the ways rural societies evolve.<br />

Far from being strictly geographically determined, they usually are the result of a deep shaping<br />

of the natural forest, <strong>and</strong> enter in complex domestication processes in order to satisfy changing<br />

rural livelihood requirements. In this sense, a better underst<strong>and</strong>ing of their characteristics <strong>and</strong><br />

functions is required for developing renewed integrated strategies for sustainable development<br />

of rural <strong>and</strong> forest management systems.<br />

References are available upon request to didier.genin@univ-provence.fr<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


D. Geni et al. 2010. A framework for characterizing convergence <strong>and</strong> discrepancy in rural forest management<br />

723<br />

Figure 1: Representation of the 11 case studies of rural forests <strong>and</strong> some significant modalities<br />

of variables on the F1xF2 plan of the Multiple Correspondences Analysis (MCA).<br />

Figure 2: Cluster Analysis of the eleven case studies of rural forests<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ouin et al. 2010. Do wooded elements in agricultural l<strong>and</strong>scape contribute to biological control in crops<br />

724<br />

Do wooded elements in agricultural l<strong>and</strong>scape contribute to biological<br />

control in crops<br />

A. Ouin 1 , M. Deconchat 2 , P. Menozzi 3 , C. Monteil 1 , L. Raison 2 , A. Roume 2 , J.P.<br />

Sarthou 1 , A. Vialatte 1 & G. Balent 2<br />

1 Université de Toulouse, UMR DYNAFOR, INRA / INP-ENSAT, BP 32607, 31326<br />

Castanet Tolosan, France.<br />

2 INRA, UMR DYNAFOR, INRA / INP-ENSAT, BP 52627, 31326 Castanet Tolosan,<br />

France<br />

3 CIRAD, UPR 10, 34 398 Montpellier, France<br />

Abstract<br />

Wooded l<strong>and</strong>scape elements are supposed to enhance biological control of pest thanks to natural<br />

enemies by provided with alternative preys, nectar <strong>and</strong> pollen resources, <strong>and</strong> refuges against<br />

unfavourable weather conditions. On the basis of two research studies in south western France<br />

we identified key features in woody elements contributing to biological control. The first study<br />

monitored in 14 winter wheat fields during 5 years aphids <strong>and</strong> aphidophagous hoverflies<br />

abundance in two l<strong>and</strong>scapes differing in woodlot density. L<strong>and</strong>scape with the higher (27%)<br />

wood cover sheltered higher hoverfly abundance during the early spring period (beginning of<br />

aphid pullulation), thus providing them with higher potential regulation capabilities. Afterwards,<br />

the difference decreased during the season. The second study dealt with ground beetles<br />

overwintering in woods. Taking advantage of emergence traps, we showed that many carabid<br />

species, including species controlling pest, overwinter in woodlots before colonizing fields.<br />

Within woodlot, distance from edge influenced abundance of overwintering carabids.<br />

Keywords: l<strong>and</strong>scape, hoverfly, carabid, woodlot, pest regulation<br />

1. Introduction<br />

Because most of the natural enemies of crop pests feed in fields but usually carry out other life<br />

cycle steps, like overwintering, in semi-natural habitats of farml<strong>and</strong> (Groeger, 1993), many<br />

studies focus on the important role of the latter, such as hedges, field margins, "beetle banks"<br />

<strong>and</strong> fallows, in increasing their populations <strong>and</strong> in improving their efficiency as control agents<br />

(Russel, 1989; L<strong>and</strong>is et al., 2000; Gurr et al., 2004). This is most probably because these seminatural<br />

habitats have a more buffered micro-climate, are less subject to agricultural disturbance<br />

than fields themselves <strong>and</strong> provide complementary or alternative food resources for larvae <strong>and</strong><br />

adults.<br />

A lot of studies have focused on overwintering of beneficial arthropods in field margins<br />

(Thomas et al., 1992). Fewer studies looked at woodlots as potential overwintering sites.<br />

Nevertheless, in temperate rural l<strong>and</strong>scapes, forest cover can be as high as 30% with a high<br />

proportion of small woodlots. In such situations, woodlots, which are in close contact with<br />

agriculture, are likely to play an important role as a refuge for overwintering beneficial<br />

arthropods (e.g. Sarthou et al., 2005).<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ouin et al. 2010. Do wooded elements in agricultural l<strong>and</strong>scape contribute to biological control in crops<br />

725<br />

The two studies presented in this paper tested the role of woodlots in controlling the population<br />

densities of beneficial insects at two different spatial scales. In the first study, we tested if there<br />

are more hoverflies <strong>and</strong> potential aphid control in wheat parcels surrounded by woodlots. In the<br />

second study, we estimated the overwintering density of ground beetles in different parts of a<br />

woodlot, from the edge to its core.<br />

2. Methodology<br />

The study region lies between the Garonne <strong>and</strong> Gers rivers, in south-western France (c. lat: 43°,<br />

long: 1°). This region is hilly (200-400m. alt.), dissected by south-north valleys, within a sub-<br />

Atlantic climate zone subject to both Mediterranean <strong>and</strong> montane influences. <strong>Forest</strong> covers 15%<br />

of the area <strong>and</strong> is composed of multiple small, private forest fragments (Balent & Courtiade,<br />

1992). <strong>L<strong>and</strong>scapes</strong> include a mix of cropl<strong>and</strong> (winter cereals, sunflower, rape, <strong>and</strong> maize or soya<br />

in irrigated lowl<strong>and</strong>), pastures, <strong>and</strong> small coppice woodlots. Semi-natural habitats are hedges,<br />

field margins, woodlots, fallow l<strong>and</strong>s <strong>and</strong> stream plus ditch edges.<br />

Spring abundance of hoverflies was studied in wheat fields in two l<strong>and</strong>scapes differing by wood<br />

density: 27% versus 15%. Wood density was determined using a l<strong>and</strong> cover classification based<br />

on four satellite images (Multib<strong>and</strong> mode, April 2001, July 2001, October 2001 <strong>and</strong> January<br />

2002). Aphidophagous hoverfly (larvae <strong>and</strong> eggs) <strong>and</strong> aphid abundances were recorded in<br />

twelve wheat fields (6 in each l<strong>and</strong>scape) in spring 2003 (April, May, June), then 14 were<br />

monitored from spring 2004 to spring 2007.<br />

We selected a woodlot that was representative of the site with respect to area (11 ha), vegetation<br />

composition (dominated by Quercus robur <strong>and</strong> Q. pubescens) <strong>and</strong> management (coppices with<br />

st<strong>and</strong>ing trees), <strong>and</strong> for which we had data on previous logging events. We set up 45 emergence<br />

traps in the woodlot, according to the distance from the boundary. The boundary of the woodlot<br />

was defined as the line joining the bases of the first trees (diameter at 1.3 m > 10 cm) belonging<br />

to the woodlot. We selected three separated zones of the woodlot according to the distance from<br />

its boundary: the edge zone (0 m to 3 m from the boundary, 12 traps), the center (75 m to 100 m,<br />

17 traps) <strong>and</strong> an intermediate zone (25 m to 50 m, 16 traps). To trap ground beetles, we chose<br />

the method of emergence traps because it makes it possible to estimate true densities of insects<br />

in very limited areas without the disadvantage of catching larvae as with soil <strong>and</strong> litter sampling,<br />

larvae being more difficult to identify. The traps were enclosed areas of 1.8 m² <strong>and</strong> were made<br />

of 0.5 mm mesh, looking like tents which sides were buried into the soil (5 cm depth). Each<br />

emergence trap included two recipients for insects: an upper recipient half-filled with 70%<br />

ethanol at the top of the trap to catch flying <strong>and</strong> climbing insects <strong>and</strong> a lower recipient level with<br />

the soil surface with 50% propylene-glycol (i.e. a pitfall trap inside the tent) to catch epigeous<br />

arthropods. The recipients were collected once a month from early March to late October 2008.<br />

Ground beetles were identified to species level (Jeannel, 1942). Then, species were pooled in<br />

three groups according to their habitat preference reported in other studies (Jeannel, 1942;<br />

Thiele, 1977; Thomas et al., 2002; Luff, 2002): forest species, which adults are caught mainly<br />

in wooded habitats with pitfall traps, open habitat species. <strong>and</strong> generalist species, found in both<br />

types of habitat. Variables used in the analysis were thus total densities of ground beetles per<br />

trap, species richness <strong>and</strong> densities of individuals in each of the three groups listed above, per<br />

trap.<br />

3. Result<br />

Spring abundance of hoverflies <strong>and</strong> aphids in wheat fields<br />

Overall, there were more aphids <strong>and</strong> hoverflies in the wooded l<strong>and</strong>scape than in the non wooded<br />

l<strong>and</strong>scape. When summing all the records in a spring, there is no significant difference in<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ouin et al. 2010. Do wooded elements in agricultural l<strong>and</strong>scape contribute to biological control in crops<br />

726<br />

hoverfly abundance (eggs+larvae) in wheat fields between the wooded <strong>and</strong> non wooded<br />

l<strong>and</strong>scapes. Compared with poorly wooded l<strong>and</strong>scapes, l<strong>and</strong>scapes with 27% of wood l<strong>and</strong>cover<br />

seemed to shelter higher hoverfly abundance during one of the key-periods for the aphid<br />

population dynamics (early spring), thus providing them with higher potential regulation<br />

capabilities. Afterwards, the difference between hoverfly population abundances decreased<br />

during the season (Figure 1).<br />

Density of overwintering ground beetles in a woodlot<br />

We collected a total of 2014 ground beetles during the whole trapping period, belonging to 48<br />

species. The total density of emerging ground beetles was four to five times higherin edges than<br />

in the inner areas of the woodlot <strong>and</strong> the mean number of species per trap was three times<br />

higher in edges than in the inner areas. Open habitats species overwintered only in the edges <strong>and</strong><br />

not in the inner areas while the two other groups of species overwintered in all the woodlot but<br />

with a marked preference for the edges (Figure 2).<br />

4. Discussion<br />

Our results show that woodlot edges are favourable for insect biodiversity in agricultural<br />

l<strong>and</strong>scapes, by (i) favouring abundances of aphidophagous hoverflies in early spring in fields,<br />

<strong>and</strong> (ii) sheltering both high species richness <strong>and</strong> abundances of ground beetles. Positive<br />

influence of woodlot edges on these insect populations seems to occur mainly during the winter<br />

period, probably by offering local environmental conditions favourable to their overwintering.<br />

Identifying such l<strong>and</strong>scape elements which favour overwintering of beneficial insects is of a<br />

particularly interest from the agronomic point of view of biological control of pests, for which<br />

early arrival of a natural enemy is required in spring (Honek, 1983; Tenhumberg & Poehling,<br />

1995; Corbett in Pickett & Bugg, 1998). Aphidophagous hoverflies distribution <strong>and</strong> abundance<br />

prove to be dependent both on l<strong>and</strong>scape parameters (Molthan, 1990), peculiar to forests <strong>and</strong><br />

crop mosaic respectively, which act at different periods through the year <strong>and</strong> sometimes with a<br />

lasting effect. Indeed south forest edges proved to be important l<strong>and</strong>scape elements for the<br />

overwintering of adult females of the beneficial aphidophagous hoverfly Episyrphus balteatus,<br />

particularly when spatially coupled with grassl<strong>and</strong>s <strong>and</strong> fallows rich in floral resources<br />

(Arrignon et al., 2007). As for north forest edges, they also proved to enhance populations of<br />

adult E. balteatus emerging in spring since they are preferential overwintering sites for the<br />

larvae, probably because of higher density of aphid populations in the fall (Sarthou et al., 2005).<br />

For ground beetles, forest edges appeared as a major reservoir of specific diversity (numerous<br />

species <strong>and</strong> their individuals) potentially able to control some pest species in the nearby crop<br />

fields, even the forest species, which are the biggest ones <strong>and</strong> are reported to prey on slugs<br />

(Kromp, 1999; Symondson et al., 2002). Moreover, woodlot edges sheltered ground beetles<br />

species whose adults live <strong>and</strong> exert beneficial influence in crops but are not usually found in<br />

wooded habitats. The impacts of edge management on these populations <strong>and</strong> their ability to<br />

move from forest to crop have to be investigated in the future. These results offer new insights<br />

for pest regulation through l<strong>and</strong>scape management, particularly by pointing the forest edges as<br />

other important l<strong>and</strong>scape elements to add to well-known set-asides <strong>and</strong> hedges in the design of<br />

semi-natural element network in agricultural l<strong>and</strong>scapes (Russel, 1989; L<strong>and</strong>is et al., 2000; Gurr<br />

et al., 2004). Undoubtedly, this 'l<strong>and</strong>scaping' management of agroecosystems will benefit from a<br />

better ecological services-based knowledge of other l<strong>and</strong>scape elements (natural grassl<strong>and</strong>s,<br />

fallows, ditch edges…) at different spatial scales, <strong>and</strong> a better integration of this knowledge in<br />

crop management.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ouin et al. 2010. Do wooded elements in agricultural l<strong>and</strong>scape contribute to biological control in crops<br />

727<br />

References<br />

Arrignon, F., M. Deconchat, J.-P. Sarthou, G. Balent <strong>and</strong> C. Monteil 2007. Modelling the<br />

overwintering strategy of a beneficial insect in a heterogeneous l<strong>and</strong>scape using a multiagent<br />

system. Ecological Modelling 205: 423-436.<br />

Balent, G. <strong>and</strong> Courtiade, B. 1992. Modelling bird communities / l<strong>and</strong>scape patterns<br />

relationships in a rural area of South-Western France. L<strong>and</strong>scape Ecology 6: 195-211.<br />

Corbett A., 1998. The importance of movements in the response of natural enemies to habitat<br />

manipulation. in Pickett C. H. & Bugg R. L. Enhancing biological control, University of<br />

California Press, Los Angeles, 1998, 422 p.<br />

Geiger, F., Wackers, F., Bianchi, F. J. J. A., 2009. Hibernation of predatory arthropods in seminatural<br />

habitats. Biocontrol 54, 529-535.<br />

Groeger, U., 1993. Investigations on the regulation of cereal aphid populations under the<br />

influence of the structure of agroecosystems. [German]. Agrarokologie, 6, 169 p.<br />

Gurr, M., Wratten, S. D. & Altieri, M. A., 2004. Ecological engineering for pest management:<br />

advances in habitat manipulation for arthropods. CAB International, 256 p.<br />

Honek, A., 1983. Factors affecting the distribution of larvae of aphid predators (Col.,<br />

Coccinellidae <strong>and</strong> Dipt., Syrphidae) in cereal st<strong>and</strong>s. Zeitschrift fur Angew<strong>and</strong>te<br />

Entomologie, 95(4), 336-345.<br />

Jeannel, R., 1942. Coléoptères carabiques, I <strong>and</strong> II, Faune de France edn. Fédération Française<br />

des Sociétés de Sciences Naturelles, Paris.<br />

Kromp, B., 1999. Carabid beetles in sustainable agriculture: a review on pest control efficacy,<br />

cultivation impacts <strong>and</strong> enhancement. Agriculture, Ecosystem, Environment 74, 187-228.<br />

L<strong>and</strong>is, D. A., Wratten, S. D. & Gurr, G. M., 2000. Habitat management to conserve natural<br />

enemies of arthropod pests in agriculture. Annu. Rev. Entomol., 45, 175-201.<br />

Luff, M. L., 2002. Carabid assemblage organization <strong>and</strong> species composition. In The<br />

agroecology of carabid beetles, ed. J. M. Holl<strong>and</strong>, pp. 41-79. Intercept, Andover.<br />

Molthan, J., 1990. Composition, community structure <strong>and</strong> seasonal abundance of hoverflies<br />

(Dipt., Syrphidae) on field margin biotopes in the Hersian Ried. Mitt. dtsch. Ges. allg.<br />

Angew, Ent. 7: 368-379.<br />

Pfiffner, L., Luka, H., 2000. Overwintering of arthropods in soils of arable fields <strong>and</strong> adjacent<br />

semi-natural habitats. Agriculture Ecosystems & Environment 78, 215-222.<br />

Sarthou, J. P., A. Ouin, F. Arrignon, G. Barreau <strong>and</strong> B. Bouyjou (2005). "L<strong>and</strong>scape parameters<br />

explain the distribution <strong>and</strong> abundance of Episyrphus balteatus (Diptera: Syrphidae)."<br />

European Journal of Entomology 102(3): 539-545.<br />

Sotherton, N. W., 1984. The Distribution <strong>and</strong> Abundance of Predatory Arthropods<br />

Overwintering on Farml<strong>and</strong>. Annals of Applied Biology 105, 423-429.<br />

Symondson, W. O. C., Sunderl<strong>and</strong>, K. D., Greenstone, M. H., 2002. Can generalist predators be<br />

effective biocontrol agents Annual Review of Entomology 47, 561-594.<br />

Tenhumberg, B. & Poehling, H. M., 1995. Syrphids as natural enemies of cereal aphids in<br />

Germany: aspects of their biology <strong>and</strong> efficacy in different years <strong>and</strong> regions. Agriculture<br />

Ecosystems & Environment, 52(1), 39-43.<br />

Thiele, H. U., 1977. Carabid beetles in their environments, Springer-Verlag, Berlin, New-York.<br />

Thomas, C. F. G., Holl<strong>and</strong>, J. M., Brown, N. J., 2002. The spatial distribution of carabid beetles<br />

in agricultural l<strong>and</strong>scapes. In The agroecology of carabid beetles, ed. J. M. Holl<strong>and</strong>, pp.<br />

305-344. Intercept, Andover.<br />

Thomas, M. B., Wratten, S. D. & Sotherton N. W., 1992. Creation of “Isl<strong>and</strong>” habitats in<br />

farml<strong>and</strong> to manipulate populations of beneficial arthropods: predator densities <strong>and</strong><br />

species composition. J. Appl. Ecol., 29: 524-531.<br />

Tscharntke T. & Kruess, A., 1999. Habitat fragmentation <strong>and</strong> biological control. In Theoretical<br />

Approaches to Biological Control , Hawkins B.A. & Cornell H.V. (Eds), Cambridge<br />

University Press, Cambridge, 190-205.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


A. Ouin et al. 2010. Do wooded elements in agricultural l<strong>and</strong>scape contribute to biological control in crops<br />

728<br />

60<br />

early average abundance of<br />

hoverflies (eggs+larvae)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

2003 2004 2005 2006 2007<br />

Year<br />

Wooded l<strong>and</strong>scape<br />

Less wooded l<strong>and</strong>scape<br />

Figure 1: Early average abundance (the first four records) of hoverflies in wheat fields in wooded <strong>and</strong> less<br />

wooded l<strong>and</strong>scapes during the five years.<br />

<strong>Forest</strong> species (n=707, s=3)<br />

Generalist species (n=710, s=12)<br />

Open habitat species (n=324, s=21)<br />

0 10 20 30 40 50<br />

0 10 20 30 40 50<br />

0 10 20 30 40 50<br />

Edge Intermediate Center<br />

Edge Intermediate Center<br />

Edge Intermediate Center<br />

Figure 2 : Density of overwintering ground beetles (number of individuals per square meter) in the edge,<br />

intermediate area <strong>and</strong> center of a woodlot, according to habitat preference of the adults. The extent of the<br />

boxes represents the first <strong>and</strong> third quantiles, the bold trait is the median.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


From ab<strong>and</strong>oned farml<strong>and</strong> to self-sustaining forests:<br />

challenges <strong>and</strong> solution


C. Estreguil & G. Caudullo 2010. Harmonized measurements of spatial pattern <strong>and</strong> connectivity<br />

730<br />

Harmonized measurements of spatial pattern <strong>and</strong> connectivity:<br />

application to forest habitats in the EBONE European Project<br />

Christine Estreguil * & Giovanni Caudullo<br />

Joint Research Centre of the European Commission<br />

Institute for Environment <strong>and</strong> Sustainability<br />

T.P.261, Via Enrico Fermi 1, 21020 Ispra (VA), Italy<br />

Abstract<br />

Within the EBONE European project (“European Biodiversity Observation NEtwork”), finegrained<br />

maps of harmonized “General Habitat Categories” are available for sixty 1 km 2 samples<br />

located in Austria, Sweden <strong>and</strong> France. Three methods were proposed to map <strong>and</strong> assess<br />

automatically the spatial pattern <strong>and</strong> connectivity of habitats. They were demonstrated for forest<br />

phanerophytes habitats. <strong>Forest</strong> spatial pattern maps were obtained from mathematical<br />

morphology (GUIDOS freeware applying a 25 m edge size) to discriminate core forest, their<br />

boundaries, connectors between core areas <strong>and</strong> islets as small non-core elements. L<strong>and</strong>scape<br />

pattern mosaic maps were generated with a L<strong>and</strong>scape Mosaic Index to characterize the forest<br />

surroundings in a disk of 25 m radius. The two pattern maps were overlaid. A “Similarity”<br />

index was proposed to assess the pre-dominance of natural habitats (thus a similar/permeable<br />

forest – non forest interface) <strong>and</strong> of anthropogenic habitats (possibly fragmentation due to<br />

cultivated or artificial l<strong>and</strong> use) in the context of the forest boundaries, connectors <strong>and</strong> islets.<br />

<strong>Forest</strong> interior areas were delineated with edge sizes depending on the similarity with their<br />

adjacent habitats. Finally, two forest connectivity indices (one with CONEFOR freeware) were<br />

computed for species with 500m dispersal capabilities on the basis of habitat availability, matrix<br />

permeability <strong>and</strong> inter-patch least-cost distances. The two indices were compared.<br />

Keywords: spatial pattern, connectivity, forest habitat, European reporting<br />

1. Introduction<br />

The European EBONE project (European Biodiversity Observation Network,<br />

http://www.ebone.wur.nl/UK) aims at European-wide habitat mapping, the delivery of habitat<br />

area estimates <strong>and</strong> the characterization of l<strong>and</strong>scape level habitat pattern, fragmentation <strong>and</strong><br />

connectivity as it is requested in the SEBI 2010 process (Streamlining European 2010<br />

Biodiversity Indicators). Methodologies should be st<strong>and</strong>ardized <strong>and</strong> easily repeatable across<br />

scales, using existing capabilities from national/regional habitat monitoring programmes.<br />

Reporting is expected using the thirteen environmental zones from the European Environmental<br />

Stratification (Metzger et al., 2005) based on climatic <strong>and</strong> topographic data at a 1 km 2 resolution.<br />

The EBONE in-situ database offers harmonized habitat field based maps (seamless vector layer<br />

with 400 m 2 Minimum Mapping Unit) over several 1km 2 samples thanks to the currently ongoing<br />

conversion of national data into the common BioHab General Habitat Categories (GHCs,<br />

Bunce et al., 2005 - figure 1). GHCs are organized in 5 super-categories i.e. whether the l<strong>and</strong><br />

surface element is ‘Urban’, ‘Cultivated’, ‘Sparsely Vegetated’ (vegetation cover below 30%),<br />

* Corresponding author. Tel.: +39 0332 785422 - Fax: +39 0332 786165<br />

Email address: christine.estreguil@jrc.ec.europa.eu<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Estreguil & G. Caudullo 2010. Harmonized measurements of spatial pattern <strong>and</strong> connectivity<br />

731<br />

‘Herbaceous’, ‘Trees or Shrubs’. Each element is described according to 16 life forms based on<br />

plant structural characteristics like plant height <strong>and</strong> leaf retention division. For example,<br />

phanerophytes are classified as forest when above 5 m height.<br />

For this study, available samples including forest phanerophytes were 16 in Sweden (NILS:<br />

National Inventory of <strong>L<strong>and</strong>scapes</strong> http://nils.slu.se), 39 in Austria (SINUS: Spatial INdices for<br />

l<strong>and</strong>-Use Sustainability), <strong>and</strong> 11 in the French Provence Cote d’Azur (PACA) region (Figure1).<br />

DESCRIPTION<br />

CODES<br />

Austria<br />

France<br />

Sweden<br />

URBAN<br />

Artificial (buildings) URB/ART X X<br />

Non vegetated URB/NON X X<br />

Vegetable gardens URB/VEG X X X<br />

Herbaceous (garden, parks) URB/GRA X X<br />

Woddy (garden tree/shrubs) URB/TRE X<br />

CULTIVATED<br />

Herbaceous crops CUL/CRO X X X<br />

Bare ground CUL/SPA X<br />

Woody crops CUL/WOC X X<br />

HERBACEOUS<br />

Leafy Hemicryptophytes HER/LHE X X X<br />

Caespitose Hemicryptophytes HER/CHE X X<br />

Cryptogams HER/CRY X<br />

Helophytes HER/HEL X<br />

Therophytes HER/THE X<br />

TREES/SHRUBS<br />

Shrubby chamaephytes TRS/SCH X<br />

Low Phanerophytes evergreen TRS/LPH X<br />

Mid Phanerophytes TRS/MPH X X<br />

Tall Phanerophytes TRS/TPH X X<br />

<strong>Forest</strong> Phanerophytes TRS/FPH X X X<br />

SPARSELY VEGETATED<br />

Aquatic SPV/AQU X X<br />

Terrestrial SPV/TER X X<br />

UNCLASSIFIED INA X<br />

Figure 1: General Habitat Categories from the available 1km 2 samples per country (left) <strong>and</strong> localization<br />

of the samples (red dots) per environmental zones (right)<br />

Available definitions of habitat pattern are first based on l<strong>and</strong>scape structure <strong>and</strong> further refined<br />

by considering organisms’ behavioral responses to the l<strong>and</strong>scape:<br />

1. The l<strong>and</strong>scape level spatial pattern of a habitat simply refers to the spatial arrangement<br />

or configuration of this habitat across the l<strong>and</strong>scape.<br />

2. Fragmentation refers to the entire process of habitat loss <strong>and</strong> isolation. Isolation means<br />

lack of connectivity <strong>and</strong> is more complicated than simple distance.<br />

3. Connectivity refers to the “degree to which the l<strong>and</strong>scape facilitates or impedes<br />

movement of organisms among resource patches”. It depends on habitat availability<br />

<strong>and</strong> spatial distribution, species’ dispersal abilities <strong>and</strong> response to the nature of the<br />

matrix.<br />

From literature on forest fragmentation, interior forest habitats are remnant minus an edge of a<br />

certain width. They retain similar abiotic <strong>and</strong> biotic conditions to pre-fragmented conditions <strong>and</strong><br />

do not experience strong influences from neighboring patches of other l<strong>and</strong> cover categories.<br />

The width of recently exposed edges, measured by two tree heights, could range from 20 m to<br />

160 m. Adjacent l<strong>and</strong> cover types possibly influence the development of the forest edge<br />

communities <strong>and</strong> interior habitat. Depending on their similarity to the forest habitats, interfaces<br />

are more or less permeable. In temperate regions, shift in l<strong>and</strong> uses at forest edges may be more<br />

important than direct forest loss. <strong>Forest</strong>s fragmented by anthropogenic sources are intuitively<br />

more vulnerable to further fragmentation than forest fragmented by natural causes.<br />

2. Methodology<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Estreguil & G. Caudullo 2010. Harmonized measurements of spatial pattern <strong>and</strong> connectivity<br />

732<br />

Three available methods are tested to characterize spatial pattern <strong>and</strong> functional connectivity for<br />

a focal habitat class <strong>and</strong> demonstrated for the focal forest phanerophytes (FPH) habitat class.<br />

2.1 Morphological Spatial Pattern Analysis (MSPA)<br />

The spatial pattern of a focal habitat class can be automatically characterized <strong>and</strong> mapped at<br />

pixel level thanks to mathematical morphology using the freeware called GUIDOS (Soille <strong>and</strong><br />

Vogt, 2009). Seven mutually exclusive pattern classes are obtained by segmenting a binary<br />

raster map (1: foreground/focal class <strong>and</strong> 0: background):<br />

1. ‘Core’: foreground pixels beyond a distance of a given size s to the background; s is the only<br />

entry parameter of the method; the input map is eroded with a Euclidian disk of radius equal to s.<br />

2. ‘Islet’: foreground pixels that do not contain any core.<br />

3. Boundary ‘Edge of core’: outer boundary pixels of a cluster of core pixels.<br />

4. Boundary ‘Edge of perforation’: inner boundary pixels of a cluster of core pixels when<br />

perforated by background pixels (like ‘holes’ inside a foreground region)<br />

5. Boundary ‘Branch’: foreground pixels with no core that is connected at one end only to a<br />

connector, an edge of core or an edge of perforation.<br />

6,7. ‘Connector’: foreground pixels with no core that connects at least two different core units<br />

(bridge) or connects to the same core unit (loop).<br />

2.2 L<strong>and</strong>scape mosaic index<br />

The l<strong>and</strong>scape context of a focal habitat class can be characterized in a Geographic Information<br />

System by applying a l<strong>and</strong>scape mosaic index (Riitters et al., 2009) on a 3-dimensional raster<br />

input map (for example, natural, agricultural <strong>and</strong> urban). L<strong>and</strong>scape pattern types are defined by<br />

placing a "window" on each pixel of the input map, calculating the proportion of the three<br />

classes within the window, <strong>and</strong> putting the result on a new map at the same location. This new<br />

map has fifteen l<strong>and</strong>scape pattern categories (see Figure 2) <strong>and</strong> the l<strong>and</strong>scape mosaic pattern<br />

map of the focal class is obtained by masking all non-focal classes. The “window” will be a<br />

Euclidian disk of radius s, like in the MSPA method, to further overlay the two pattern maps.<br />

Figure 2: the fifteen l<strong>and</strong>scape pattern types derived with the l<strong>and</strong>scape mosaic index<br />

The MSPA <strong>and</strong> the l<strong>and</strong>scape mosaic pattern maps will then be overlaid to provide the<br />

l<strong>and</strong>scape context composition in terms of mosaic pattern types for each non-core MSPA class<br />

(boundary, connector, <strong>and</strong> islet). A new “similarity” index (SI) is proposed to translate the<br />

anthropogenic or natural dominance in the surroundings. When the mosaic pattern is NN <strong>and</strong><br />

the focal class forest, the context is similarly 100% natural, possibly permeable <strong>and</strong> most<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Estreguil & G. Caudullo 2010. Harmonized measurements of spatial pattern <strong>and</strong> connectivity<br />

733<br />

probably due to natural fragmentation causes. Anthropogenic fragmentation causes in predominant<br />

natural context are pointed at by using patterns N or (Nu, Nua, Na) in the formula.<br />

( MosaicPattern)<br />

MSPAClass<br />

SI ( MosaicPattern)<br />

MSPA Class<br />

= (1)<br />

MSPAClass<br />

“Interior” areas are delineated as core areas plus the NN part of the MSPA boundary edge.<br />

2.3 Connectivity assessment<br />

The Probability of Connectivity (PC) index for a focal class, calculated with the software<br />

Conefor Sensinode (Saura <strong>and</strong> Torne, 2009 at http://www.conefor.org), is based on topology<br />

(inter-patch distances), patch attributes like area <strong>and</strong> species specific dispersal ability. PC will<br />

be processed with the probability of dispersal, being a decreasing exponential function of the<br />

effective distance, matching to a 50% probability for a specific average dispersal distance. The<br />

effective distance is a value of movement cost through different habitats that is obtained through<br />

least-cost path algorithms, thus considering the l<strong>and</strong>scape permeability between the focal<br />

patches. PC has a bounded range of variation from 0 to 1. The cost distance matching the 50%<br />

probability (p = 0.5, cost d50% ) corresponds to the average dispersal distance (d 50% ) multiplied by<br />

the average friction per distance unit (avg_f). The average friction is set at half a logarithmic<br />

scale of frictions, being from 1 to 10,000 (avg_f = 100). PC is made comparable to the available<br />

habitat in the total l<strong>and</strong>scape area, by computing its square root (RPC).<br />

n<br />

n<br />

∑∑<br />

i= 1 j=<br />

1<br />

a ⋅ a ⋅ p<br />

i<br />

PC = (2)<br />

A<br />

2<br />

L<br />

RPC= PC (2bis)<br />

j<br />

ij<br />

with:<br />

a i a j = area of patches<br />

ln(0.5)<br />

A L = total l<strong>and</strong>scape area<br />

k =<br />

cost<br />

p ij = e k • costij d 50%<br />

cost d50% = avg_f • d 50%<br />

Another index, adapted from Hanski (1994), called Isolation Sensitive Index (IsoSi) is proposed.<br />

It is similar to PC but accounts for solely the arrival patch area size for each pair of patches. The<br />

l<strong>and</strong>scape area (A L ) <strong>and</strong> the number of links (node to node) are used for normalization purposes.<br />

IsoSi<br />

=<br />

A<br />

n<br />

∑<br />

L<br />

a ⋅ p<br />

i<br />

i≠<br />

j,i=<br />

1<br />

ij<br />

⋅ (n−1)<br />

(3)<br />

with:<br />

n = number of patches (nodes)<br />

3. Result <strong>and</strong> discussion<br />

3.1 Pattern characterization based on MSPA <strong>and</strong> l<strong>and</strong>scape mosaic index<br />

First, the GHCs vector maps of all available samples (figure 1) were rasterised at 1 m spatial<br />

resolution, re-classified into forest phanerophytes (FPH)-non forest <strong>and</strong> processed with<br />

GUIDOS using a narrow forest edge width (s equal to 25 m). The local morphology of the FPH<br />

habitat cover was mapped according to 4 main pattern classes (upper figure 3) <strong>and</strong> their forest<br />

area share (figure 4 left) was calculated: core, boundary (edges of core, perforation <strong>and</strong> branch),<br />

connector (bridge <strong>and</strong> loop) <strong>and</strong> islet.<br />

Second, the GHCs 1 m raster maps were re-classified into natural (TRS, HER, SPV), cultivated<br />

(CUL) <strong>and</strong> urban (URB) habitat types (figure 1). The fifteen l<strong>and</strong>scape pattern types were<br />

mapped by applying the mosaic index using a 25 m radius disk. The non-FPH classes were<br />

masked. The l<strong>and</strong>scape context map of FPH habitats enables to visualize <strong>and</strong> characterize FPH<br />

interface zones (NN discriminated from Nu for example) (figure 3 bottom), <strong>and</strong> compute forest<br />

proportion of the 4 main l<strong>and</strong>scape FPH pattern types for each available sample (figure 4, right):<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Estreguil & G. Caudullo 2010. Harmonized measurements of spatial pattern <strong>and</strong> connectivity<br />

734<br />

- Two natural forest l<strong>and</strong>scape patterns where FPH habitats have no (NN) or not significant (N)<br />

edge shared with cultivated <strong>and</strong>/or artificial habitats, interface zones are possibly permeable.<br />

- Mixed natural forest l<strong>and</strong>scape pattern (Nu, Nua, Na) where FPH habitats have possibly less<br />

permeable interfaces as being adjacent with cultivated <strong>and</strong> urban types of habitats<br />

- “Some natural” forest l<strong>and</strong>scape (all others) where FPH habitats are pre-dominantly embedded<br />

in non-natural context of cultivated <strong>and</strong> urban types of habitats.<br />

Figure 3: Example of two samples in Austria: <strong>Forest</strong> MSPA (upper) <strong>and</strong> l<strong>and</strong>scape mosaic (bottom) maps<br />

MSPA<br />

Core<br />

Connector<br />

Boundary<br />

Islet<br />

Mosaic<br />

Natural NN<br />

Mixed natural<br />

Natural N<br />

Some natural<br />

100%<br />

100%<br />

90%<br />

90%<br />

<strong>Forest</strong> proportion of MSPA classes<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

<strong>Forest</strong> proportion of Mosaic classes<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

0%<br />

au113<br />

au331<br />

au113<br />

au331<br />

Figure 4: MSPA <strong>and</strong> l<strong>and</strong>scape mosaic class proportion for the same two samples in Austria<br />

The MSPA <strong>and</strong> the mosaic pattern maps were overlaid to compute the Similarity Index for noncore<br />

MSPA classes, <strong>and</strong> delineate “interior” forest areas which edge width depends on adjacent<br />

habitats. <strong>Forest</strong> proportion of “interior” <strong>and</strong> core areas can be compared in table 1 for two<br />

samples with different permeable boundary contexts as illustrated by the proportion of NN in<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


C. Estreguil & G. Caudullo 2010. Harmonized measurements of spatial pattern <strong>and</strong> connectivity<br />

735<br />

the boundary MSPA class (SI (NN) Boundary ). Also, boundaries are more exposed to anthropogenic<br />

fragmentation in pre-dominant natural context (SI(NuaNaNu) Boundary ) in the less permeable Au113.<br />

Table 1: <strong>Forest</strong> (FPH) proportion of “interior” <strong>and</strong> Core area, <strong>and</strong> the Similarity Index applied to<br />

boundaries in two samples in Austria (Continental zone).<br />

Samples Id Core FPH Interior FPH SI (NN) Boundary SI (NuaNaNu) Boundary SI (some nat.) Boundary<br />

Au113 40.1% 43.5% 18.6% 36.7% 6.4%<br />

Au331 26.5% 41.3% 58.9% 13.7% 3.6%<br />

3.2 Connectivity<br />

Connectivity indices PC, RPC <strong>and</strong> IsoSi were calculated for species dispersing at 500 m average<br />

dispersal distance. Costs of movement (friction) were assigned to every habitat types using a<br />

logarithmic increment values from FPH (lowest friction 1) to urban habitats (highest friction<br />

10.000). The parameter cost d50% was 50.000. RPC <strong>and</strong> IsoSi behaved differently (see Table 2<br />

with the sample Au113 with fewer nodes <strong>and</strong> a less permeable context than the sample Au331).<br />

Table 2: Connectivity indices in two different spatial configurations <strong>and</strong> permeability contexts<br />

Samples Id FPH area % Nodes number PC RPC IsoSi<br />

Au113 63 % 33 35 % 59 % 50%<br />

Au331 57 % 85 31 % 56 % 55%<br />

IsoSi is more sensitive to the inter-patch l<strong>and</strong>scape matrix permeability, possible barrier effects<br />

<strong>and</strong> is thus more focused on the probability of species movement. This was expected since the<br />

weight for areas (intra-patch) is the same than p ij while it is double in the PC (<strong>and</strong> RPC) index.<br />

In contrast, PC (<strong>and</strong> RPC) reacts better to habitat availability (its intra <strong>and</strong> inter-connectivity).<br />

Are the two indices necessary to correctly describe l<strong>and</strong>scape connectivity <strong>and</strong> its permeability<br />

How sensitive a connectivity index should be to the matrix permeability More research needed.<br />

4. Conclusions<br />

For all sample, harmonized pattern <strong>and</strong> connectivity maps <strong>and</strong> tabular data were organized per<br />

environmental zone. They will be incorporated into the EBONE data management structure<br />

prototype to be ready at the end of the project (2012). The methods here proposed are currently<br />

repeated over the available Earth Observation based l<strong>and</strong> cover maps to prepare the integration<br />

of EO based <strong>and</strong> in situ habitat pattern assessment in the view of extending the geographical<br />

extent of habitat pattern/connectivity information available for biodiversity assessment<br />

(Estreguil <strong>and</strong> Mouton 2009).<br />

References<br />

Bunce, B. et al, 2005. A st<strong>and</strong>ardized procedure for surveillance <strong>and</strong> monitoring European<br />

habitats <strong>and</strong> provision of spatial data. L<strong>and</strong>scape Ecology, 23:11–25<br />

Estreguil C. <strong>and</strong> Mouton C., 2009. Measuring <strong>and</strong> reporting on forest l<strong>and</strong>scape pattern,<br />

fragmentation <strong>and</strong> connectivity in Europe: methodology <strong>and</strong> indicator layers. Office for<br />

Official Publications of the European Communities, EUR23841EN.<br />

Metzger M.J., et al 2005. A climatic stratification of the environment of Europe. <strong>Global</strong><br />

Ecology <strong>and</strong> Biogeography. 14: 549-563.<br />

Riitters K.H., Wickham J.D. <strong>and</strong> Wade T.G., 2009. An indicator of forest dynamics using a<br />

shifting l<strong>and</strong>scape mosaic. Ecological Indicators 9:107-117<br />

Saura, S., Torne´J. , 2009. Conefor Sensinode 2.2: A software package for quantifying the<br />

importance of habitat patches for l<strong>and</strong>scape connectivity, Environ. Model. Softw. (2008),<br />

doi:10.1016/j.envsoft.2008.05.005.<br />

Soille <strong>and</strong> Vogt, 2009. Morphological segmentation of binary patterns. Patterns Recognition<br />

Letters. doi:10.1016/j.patrec.2008.10.015<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.M. Rey Benayas 2010. Restoration of biodiversity <strong>and</strong> ecosystem services in cropl<strong>and</strong><br />

736<br />

Restoration of biodiversity <strong>and</strong> ecosystem services in cropl<strong>and</strong>.<br />

Further research is needed but action is desperately needed<br />

José M. Rey Benayas *<br />

Dpto. de Ecología, Universidad de Alcalá, Spain<br />

Abstract<br />

Farml<strong>and</strong> currently extends on more than 40% of the l<strong>and</strong>’s surface. Secondary succession on<br />

ab<strong>and</strong>oned agricultural l<strong>and</strong> is often slow owing to biotic <strong>and</strong> abiotic limitations as it occurs in<br />

e.g. Mediterranean environments. Tree plantations can be expensive if large areas are to be<br />

restored <strong>and</strong> they are often controversial because they negatively impact on species of<br />

conservation concern. We suggest “woodl<strong>and</strong> islets” as an alternative approach to designing<br />

ecological restoration in extensive agricultural l<strong>and</strong>scapes. This approach allows conciliate<br />

farml<strong>and</strong> production, conservation of values linked to cultural l<strong>and</strong>scapes, enhancement of<br />

biodiversity <strong>and</strong> provision of a range of ecosystem services. If “further research is needed”,<br />

“action is desperately needed”. The International Foundation for Ecosystem Restoration is<br />

developing the “Islets <strong>and</strong> coasts in agricultural seas” Initiative to implement demonstration<br />

projects of conciliation of farml<strong>and</strong> production <strong>and</strong> enhancement of biodiversity <strong>and</strong> ecosystem<br />

services. The implemented restoration actions intended to “renaturalize” agricultural l<strong>and</strong>scapes<br />

are accompanied by a variety of social <strong>and</strong> educational values including citizen science.<br />

Keywords: conciliation, ecological <strong>and</strong> societal benefits, secondary succession, tree plantations,<br />

woodl<strong>and</strong> islets<br />

1. Introduction<br />

Human cultural evolution has left an unprecedented ecological footprint on Earth, so as to affect<br />

currently about 80% of the planet’s surface. This implies large losses of biodiversity <strong>and</strong> of the<br />

variety, amount <strong>and</strong> quality of ecosystem services, which compromises the sustainability of ecosociological<br />

systems (Ostrom 2009). Unfortunately, predictions suggest that the ecological<br />

human footprint will exp<strong>and</strong> in the future (Hockley et al. 2008). A large part of global<br />

environmental degradation is due to the expansion of the agricultural <strong>and</strong> livestock boundary in<br />

most parts of the world together with agricultural intensification. Currently, agricultural l<strong>and</strong>s<br />

<strong>and</strong> pasturel<strong>and</strong>s make up over 40% of the terrestrial surface, in detriment of natural vegetation<br />

cover (Foley et al. 2005). On the contrary, large extents of agricultural l<strong>and</strong> <strong>and</strong> pastures have<br />

been ab<strong>and</strong>oned over the last few decades for economic reasons. Any event that spurs an either<br />

“positive” or “negative” effect in these systems can have a relevant ecological <strong>and</strong> socioeconomic<br />

impact.<br />

Ecological restoration aims to recover, at least partially, the characteristics of an ecosystem,<br />

such as its biodiversity <strong>and</strong> function, before degradation or destruction occurred, generally as a<br />

result of human activities (Rey Benayas et al. 2009). Restoration ecology actions have been<br />

increasingly implemented, supported by agreements by global politicians such as the<br />

* Corresponding author. Tel.: +34 91 885 4987 - Fax: +34 91 885 4929<br />

Email address: josem.rey@uah.es<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.M. Rey Benayas 2010. Restoration of biodiversity <strong>and</strong> ecosystem services in cropl<strong>and</strong><br />

737<br />

Convention for Biological Diversity <strong>and</strong> the sustained global biodiversity crisis (Sutherl<strong>and</strong> et al.<br />

2009). The scientific community must search for ecological restoration protocols <strong>and</strong> models<br />

that allow a synergy between the exploitation of ecosystems <strong>and</strong> nature conservation, which will<br />

in turn improve the sustainability of systems for the exploitation of natural resources.<br />

2. The agriculture <strong>and</strong> conservation paradox<br />

Few human activities are as paradoxical as agriculture in terms of their role for nature<br />

conservation. On one side, agricultural activities are the main cause of deforestation worldwide,<br />

which has occurred at an estimated global rate of 130,000 km 2 per year over the last five years<br />

(FAO 2006). Traditional agriculture allowed remnants of natural vegetation to remain on steep<br />

hillsides, valleys, rocky outcrops, shallow soils, saline <strong>and</strong> infertile areas, property boundaries<br />

<strong>and</strong> track edges. In recent history farming practices in many areas have become intensified, <strong>and</strong><br />

increasing amounts of water, fuel, fertilizers, pesticides, <strong>and</strong> herbicides are used worldwide to<br />

increase food <strong>and</strong> fiber production. E.g., global area equipped for irrigation exp<strong>and</strong>ed by 0.3%<br />

to 280 million ha between 2004 <strong>and</strong> 2005, <strong>and</strong> at present irrigated area accounts for ca. 20% of<br />

cultivated l<strong>and</strong>. Intensification of l<strong>and</strong> use has brought remnant areas of natural vegetation into<br />

mainstream agriculture <strong>and</strong> many such areas have been lost or severely degraded. The<br />

conversion of natural ecosystems to human l<strong>and</strong>-uses seems to have ensured our food supplies<br />

at a global scale. However, food security has damaged the regulation function of ecosystems.<br />

Whereas the provision of environmental services such as crops <strong>and</strong> livestock production have<br />

increased, hydrological <strong>and</strong> climate regulation, soil retention, <strong>and</strong> greenhouse gas mitigation<br />

have decreased as a consequence of overall degradation of ecosystem services by 60% in the<br />

last 50 years (Millenium Ecosystem Assessment 2005).<br />

However, deforested habitats as a result of agricultural activities have been often granted a<br />

relevant role in nature conservation (Kleijn et al. 2006). Thus, of the seven main categories of<br />

terrestrial habitats in the EU Habitats Directive, four include agricultural <strong>and</strong> livestock l<strong>and</strong> uses<br />

(pastures, scrub, “dehesas” <strong>and</strong> “montados”, etc.). Several species, some of them endangered,<br />

especially birds, depend on agrarian systems (http://www.birdlife.org/). Agricultural<br />

intensification can have a negative impact on these values, but so can agricultural<br />

ab<strong>and</strong>onment <strong>and</strong>, particularly, afforestation of former cropl<strong>and</strong>. It seems that<br />

agriculture, woodl<strong>and</strong>, <strong>and</strong> biological conservation are in a permanent <strong>and</strong> irreconcilable<br />

conflict, the agriculture <strong>and</strong> conservation paradox. This creates a dilemma in woodl<strong>and</strong><br />

restoration projects, which can only be resolved by considering the relative values<br />

associated with woodl<strong>and</strong> vs. agricultural ecosystems. Can we solve this dilemma<br />

3. Contrasting approaches for vegetation restoration in ab<strong>and</strong>oned cropl<strong>and</strong><br />

Ab<strong>and</strong>oned agricultural l<strong>and</strong>, i.e. cropl<strong>and</strong> <strong>and</strong> pastures where extensive livestock farming has<br />

been removed, can be subject of secondary succession or passive vegetation restoration.<br />

Worldwide, l<strong>and</strong> ab<strong>and</strong>onment <strong>and</strong> passive restoration have revegetated a greater surface area<br />

<strong>and</strong> at a smaller cost than active restoration (45,000 Km 2 /year vs. 28,000 Km 2 /year, respectively,<br />

in 2000-2005; in Europe, agricultural l<strong>and</strong> area have declined by ca. 13% between 1961 <strong>and</strong><br />

2000). Passive restoration is cheap <strong>and</strong> genuine since all ecological filters are at play, It is<br />

generally fast in productive environments, but usually slow in low productivity environments<br />

such as the Mediterranean. This is because woody vegetation establishment is limited, since<br />

conditions in bare areas are different to those of places where trees <strong>and</strong> shrubs regenerate<br />

naturally (Rey Benayas et al. 2005). A key bottle-neck that hinders revegetation is lack of<br />

propagules due to absence of mother trees <strong>and</strong> shrubs due to complete destruction of the original<br />

vegetation.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.M. Rey Benayas 2010. Restoration of biodiversity <strong>and</strong> ecosystem services in cropl<strong>and</strong><br />

738<br />

Active restoration basically consists in planting trees <strong>and</strong> shrubs. It is needed when ab<strong>and</strong>oned<br />

l<strong>and</strong> suffers continuos degradation, local vegetation cover cannot be recovered <strong>and</strong> secondary<br />

succession should be accelerated, among others. The European Union has paid farmers for<br />

afforestation since 1990, offering grants to turn farml<strong>and</strong> back into forest <strong>and</strong> payments for the<br />

management of such forest. Between 1993 <strong>and</strong> 1997, EU afforestation policies made possible<br />

the afforestation of over 5,000 Km 2 of l<strong>and</strong>. A second program, running between 2000 <strong>and</strong> 2006,<br />

afforested in excess of ca. 1,000 Km 2 of l<strong>and</strong>. A third such program began in 2007. Much of the<br />

forthcoming afforested l<strong>and</strong> will occur in former vineyards that occupied ca. 3.4 million<br />

hectares in the EU-25 by 2006. In Spain, during the first five years of the 2000 decade, the<br />

recovery of forests has reached 152,400 ha/year, of which 16% are plantations (FAO 2006). On<br />

top of this, 700,000 ha were reforested during 1993-2006 thanks to the incentives of the<br />

Common Agricultural Policy of the European Union. This amount of reforested agricultural<br />

l<strong>and</strong> will increase as removal of vineyards is intended to be 175,000 ha in the 2008-2011 period,<br />

44,090 of which have already been extirpated in 2008-2009.<br />

Most afforestations of former cropl<strong>and</strong> in Mediterranean Europe has based on coniferous<br />

species, though other species such as eucalyptus <strong>and</strong> oaks have been used. There is debate about<br />

the ecological benefits of these afforestations, some of which are controversial. The trade-off<br />

among different types of ecosystem services <strong>and</strong> biodiversity values is at the core of such debate.<br />

For instance, the fast-growing plantations are better for carbon sequestration per time unit than<br />

secondary succession of shrubl<strong>and</strong> <strong>and</strong> woodl<strong>and</strong>, <strong>and</strong> they may provide timber in the future.<br />

However, they may cause damage to open habitat species, especially birds which are of<br />

conservation concern in Europe, by substracting amount of high quality habitat <strong>and</strong> increasing<br />

risk of predation. Further, these plantations have been shown to be suitable habitats just for<br />

generalist forest birds but not for specialist forest birds (Rey Benayas et al. 2010a).<br />

4. Innovating vegetation restoration in agricultural l<strong>and</strong>scapes<br />

The reconstruction of vegetation in a l<strong>and</strong>scape (“where <strong>and</strong> when to revegetate”) is an issue<br />

that deserves to become a research priority (Munro et al. 2009). The problems with existing<br />

methods to restore woodl<strong>and</strong>s in agricultural l<strong>and</strong>scapes should not give rise to pessimism, but<br />

instead inspire researchers to devise new creative solutions. A realistic view of conservation<br />

must acknowledge the conservation value of the agricultural matrix in forest l<strong>and</strong>scapes. A<br />

focus on the matrix is required if we are serious about solving the biodiversity crisis, <strong>and</strong> that<br />

matrix is usually an agro-ecosystem of some sort (V<strong>and</strong>ermeer <strong>and</strong> Perfecto 2007). Thus, agrosuccessional<br />

restoration schemes are proposed, which include agroecological <strong>and</strong> agroforestry<br />

techniques as a step prior to forest restoration (Vieira et al. 2009).<br />

We suggest a different concept for designing restoration of forest ecosystems on agricultural<br />

l<strong>and</strong>, which uses small-scale active restoration as a driver for passive restoration over much<br />

larger areas. This could increase the economic feasibility of large-scale restoration projects <strong>and</strong><br />

facilitate the involvement of local human communities in the restoration process. Establishment<br />

of “woodl<strong>and</strong> islets” is an approach to designing restoration of woodl<strong>and</strong>s in extensive<br />

agricultural l<strong>and</strong>scapes where no remnants of native natural vegetation exist (Rey Benayas et al.<br />

2008). Whereas passive restoration leaves reforestation to chance <strong>and</strong> active restoration usually<br />

requires a large input of resources, the woodl<strong>and</strong> islets approach provides an intermediate<br />

degree of intervention. It allows direction of secondary succession by establishing small<br />

colonisation foci, while using a fraction of the resources required for large-scale reforestation. In<br />

addition it maintains flexibility of l<strong>and</strong> use, which is critical in agricultural l<strong>and</strong>scapes where<br />

l<strong>and</strong> use is subject to a number of fluctuating policy <strong>and</strong> economic drivers. The approach<br />

involves planting a number of small (some tens or a few hundreds of m 2 ), densely-planted (e.g.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.M. Rey Benayas 2010. Restoration of biodiversity <strong>and</strong> ecosystem services in cropl<strong>and</strong><br />

739<br />

one introduced seedling per 2 m 2 ), <strong>and</strong> sparse (some tens or hundreds of m apart) blocks of<br />

native trees within agricultural l<strong>and</strong> that together occupy a tiny fraction of the area of target l<strong>and</strong><br />

to be restored, e.g.


J.M. Rey Benayas 2010. Restoration of biodiversity <strong>and</strong> ecosystem services in cropl<strong>and</strong><br />

740<br />

other benefits than production, such benefits are the primary objective of the woodl<strong>and</strong> islets<br />

approach. A key distinction is the l<strong>and</strong>scape emphasis on a planned planting of islets to<br />

maximise benefits to biodiversity, <strong>and</strong> the potential of allowing the islets to form foci for largerscale<br />

reforestation of intervening l<strong>and</strong>. Furthermore, if the surrounding l<strong>and</strong> is to be farmed, its<br />

management can be designed to make use of the ecosystem services provided by the islets.<br />

5. Action is desperately needed<br />

We have been running research on the woodl<strong>and</strong> islets approach since 1993. As researchers, we<br />

have produced a number of publications that illustrate observations, patterns, processes,<br />

hypotheses, etc. for particular case studies. In most of our publications, as it usually occurs in<br />

the scientific literature, we mention that “further research is needed” to tight up popping up<br />

issues. It is certainly our responsibility as academia professionals to research further but, in my<br />

view, we must transfer as well the created knowledge to projects in the real world.<br />

To target this aim, the International Foundation for Ecosystem Restoration<br />

(www.fundacionfire.org) is developing the “Islets <strong>and</strong> coasts in agricultural seas” Initiative to<br />

implement demonstration projects of conciliation of farml<strong>and</strong> production <strong>and</strong> enhancement of<br />

biodiversity <strong>and</strong> ecosystem services. Restoration actions in these projects include the following.<br />

• Introduction of woodl<strong>and</strong> islets of native plant species as explained above.<br />

• Revegetation of field boundaries <strong>and</strong> way sides (“coasts” in the agricultural “seas”).<br />

• Rehabilitation <strong>and</strong> construction of water spots (ponds, springs, drinking troughs), which<br />

are of critical importance for wildlife, especially for amphibians.<br />

• Installation of nest boxes for birds, particularly those that are useful to farmers<br />

(insectivorous birds, raptors). The young pine plantations on afforested former cropl<strong>and</strong><br />

are suitable habitats for this action.<br />

• Installation of perches for birds in ab<strong>and</strong>oned cropl<strong>and</strong> to enhance seed rain on them.<br />

• Rehabilitation <strong>and</strong> construction of stone mounds <strong>and</strong> walls, as well as constructions of<br />

rural architecture.<br />

• Propagation <strong>and</strong> plantation of singular fruit trees, i.e. old <strong>and</strong> healthy fruit trees of local<br />

varieties that are rapidly going extinct. We use the propagated trees for their plantation<br />

in restored cropl<strong>and</strong> <strong>and</strong> for creating a genetic reserve.<br />

• Plantation of olive groves for CO 2 emission compensation, where we implement the<br />

actions above.<br />

The implemented restoration actions intended to “renaturalize” agricultural l<strong>and</strong>scapes are<br />

accompanied by a variety of social <strong>and</strong> educational benefits. Most work in the field is developed<br />

by volunteers as we pursue environmental education <strong>and</strong> sensibility. These volunteers derive<br />

from environmentalist <strong>and</strong> conservationist groups, schools, <strong>and</strong> companies (corporative<br />

volunteers). We are developing a program of citizen science consisting in the monitoring of<br />

introduced plants <strong>and</strong> nest boxes. Our restoration projects are visited by students from schools,<br />

universities <strong>and</strong> specialized training courses (Rey Benayas et al. 2010b). The nest boxes are<br />

built by h<strong>and</strong>icapped local people under the supervision of professional carpenters. The<br />

propagation of singular fruit trees is done by horticulture schools as well as by professional<br />

nurseries. Some restoration actions <strong>and</strong> their maintenance are contracted to local farmers.<br />

Overall, these projects are producing income <strong>and</strong> opportunities to local communities.<br />

Importantly, l<strong>and</strong> owners must be explicitly rewarded for the restoration actions occurring in<br />

their properties. Beyond the subsidies from agri-environmental schemes, other approaches such<br />

as payment for environmental services <strong>and</strong> tax deduction must be rapid <strong>and</strong> generously<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


J.M. Rey Benayas 2010. Restoration of biodiversity <strong>and</strong> ecosystem services in cropl<strong>and</strong><br />

741<br />

implemented in a time when society dem<strong>and</strong>s from agricultural l<strong>and</strong> <strong>and</strong> farmers much more<br />

than food <strong>and</strong> fiber production.<br />

Acknowledgements. I am indebted to all co-authors in my referred publications in this paper,<br />

particularly J.M. Bullock <strong>and</strong> A.C. Newton. Projects from the Spanish Ministry of Science<br />

<strong>and</strong> Education (CGL2007-60533-BOS) <strong>and</strong> the Government of Madrid (S2009AMB-<br />

1783, REMEDINAL) are currently providing financial support of this body of research.<br />

References<br />

FAO, 2006. The <strong>Global</strong> <strong>Forest</strong> Resources Assessment 2005. FAO, Rome (URL:<br />

http://www.fao.org/forestry).<br />

Foley, J.A., DeFries, R., Asner, G.P., Barford, C., Bonan, G., Carpenter, S.R., Chapin, F. S.,<br />

Coe, M.T., Daily, G.C., Gibbs, H.K., Helkowski, J.H., Holloway, T., Howard, E.A.,<br />

Kucharik, C.J., Monfreda, C., Patz, J.A.; Prentice, I.C,.Ramankutty, N. <strong>and</strong> Snyder, P.K.,<br />

2005. <strong>Global</strong> consequences of l<strong>and</strong> use. Science, 309: 570-574.<br />

Hockley, N.J., Jones, J.P.G. <strong>and</strong> Gibbons, J. 2008 Technological progress must accelerate to<br />

reduce ecological footprint overshoot. Frontiers in Ecology <strong>and</strong> the Environment, 6: 122-123.<br />

Kleijn, D., Baquero, R.A., Clough, Y., Díaz, M., Esteban, J., Fernández, F., Gabriel, D., Herzog,<br />

F., Holzschuh, A., Jöhl, R., Knop, E., Kruess, A., Marshall, E.J., Steffan-Dewenter, I.,<br />

Tscharntke, T., Verhulst, J., West, T.M. <strong>and</strong> Yela, J.L. 2006. Mixed biodiversity benefits of<br />

agro-environment schemes in five European countries. Ecology Letters, 9: 243-254.<br />

Millenium Ecosystem Assessment (Ed.), 2005. Ecosystems <strong>and</strong> Human-Well Being. Isl<strong>and</strong><br />

Press, New York.<br />

Munro, N. T., Fischer, J., Wood, J. <strong>and</strong> Lindenmayer, D.B., 2009. Revegetation in agricultural<br />

areas: the development of structural complexity <strong>and</strong> floristic diversity. Ecological<br />

Applications, 19: 1197-2010.<br />

Ostrom, E., 2009. A general framework for analyzing sustainability of social-ecological systems.<br />

Science, 325: 419-422.<br />

Rey Benayas, J.M., Navarro, J., Espigares, T., Nicolau, J.M. <strong>and</strong> Zavala, M.A. 2005. Effects of<br />

artificial shading <strong>and</strong> weed mowing on reforestation of Mediterranean ab<strong>and</strong>oned cropl<strong>and</strong><br />

with contrasting Quercus species. <strong>Forest</strong> Ecology & Management, 212: 302-314.<br />

Rey Benayas, J.M., Bullock, J.M. <strong>and</strong> Newton, A.C. 2008. Creating woodl<strong>and</strong> islets to reconcile<br />

ecological restoration, conservation, <strong>and</strong> agricultural l<strong>and</strong> use. Frontiers in Ecology <strong>and</strong> the<br />

Environment, 6: 329-33.<br />

Rey Benayas, J.M., Newton, A.C., Diaz, A. <strong>and</strong> Bullock, J.M., 2009. Enhancement of<br />

biodiversity <strong>and</strong> ecosystem services by ecological restoration: a meta-analysis. Science, 325:<br />

1121-1124.<br />

Rey Benayas, J.M., Galván, I. <strong>and</strong> Carrascal, L.M., 2010a. Differential effects of vegetation<br />

restoration in Mediterranean ab<strong>and</strong>oned cropl<strong>and</strong> by secondary succession <strong>and</strong> pine<br />

plantations on bird assemblages. <strong>Forest</strong> Ecology <strong>and</strong> Management,<br />

doi:10.1016/j.foreco.2010.04.004.<br />

Rey Benayas, J.M., Escudero, A., Martín Duque, J.F., Nicolau, J.M., Villar, P., García de Jalón,<br />

D. <strong>and</strong> Balaguer, L., 2010b. A multi-institutional Spanish Master in Ecosystem Restoration:<br />

vision <strong>and</strong> four-year experience. Ecological Restoration, 28: 188-192.<br />

Sutherl<strong>and</strong>, W.J. et al. (44 authors), 2009. One hundred questions of importance to the<br />

conservation of global biological diversity. Conservation Biology, 23: 557–567.<br />

V<strong>and</strong>ermeer, J. <strong>and</strong> Perfecto, I., 2007. The agricultural matrix <strong>and</strong> a future paradigm for<br />

conservation. Conservation Biology, 21: 274-277.<br />

Vieira, D.L.M., Holl, K.D. <strong>and</strong> Peneireiro, F.M., 2009. Agro-successional restoration as a<br />

strategy to facilitate tropical forest recovery. Restoration Ecology, 17: 451–459.<br />

<strong>Forest</strong> <strong>L<strong>and</strong>scapes</strong> <strong>and</strong> <strong>Global</strong> <strong>Change</strong>-New Frontiers in Management, Conservation <strong>and</strong> Restoration. Proceedings of the IUFRO L<strong>and</strong>scape Ecology<br />

Working Group International Conference, September 21-27, 2010, Bragança, Portugal. J.C. Azevedo, M. Feliciano, J. Castro & M.A. Pinto (eds.)<br />

2010, Instituto Politécnico de Bragança, Bragança, Portugal.


IUFRO Unit 8.01.02 L<strong>and</strong>scape Ecology<br />

CIMO - Centro de Investigação de Montanha, Portugal<br />

IPB - Instituto Politécnico de Bragança, Portugal<br />

Sociedade Portuguesa de Ciências Florestais<br />

(Portuguese Society of <strong>Forest</strong> Sciences)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!