Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi- scale approach in the D

Sahara-Sahel

Frederico da Costa Santarém

Doutoramento em Biodiversidade, Genética e Evolução Departamento de Biologia, Faculdade de Ciências da Universidade do Porto

2020 Orientador

José Carlos Brito, Investigador Principal, Faculdade de Ciências da Universidade do Porto, Porto, Portugal Coorientador Jarkko Saarinen, Professor, University of Oulu, Oulu, Finland

Foreword

In compliance with the no. 2 of article 4 of the General Regulation of Third Cycles of the University of Porto and with the article 31 of the Decree-Law no. 74/2006 of 24 March, with the alteration introduced by the Decree-Law no. 230/2009, of 14 September, the results of already published works were totally used and included in some of the chapters of this dissertation. As these works were performed in collaboration with other authors, the candidate clarifies that, in all these works, participated in obtaining, analysing, and discussing the results, as well as in the writing of the published forms. The Faculty of Sciences of the University of Porto was the home institution of the candidate, and the work was supervised by Professor José Carlos Brito (CIBIO-InBIO – Research Center in Biodiversity and Genetic Resources, Vairão, Portugal and Faculty of Sciences of the University of Porto, Porto, Portugal) and co-supervised by Professor Jarkko Saarinen (Geography Research Unit of the University of Oulu, Oulu, Finland, and Distinguished Visiting Professor (Sustainability management) in the University of Johannesburg, South ). The work was developed at CIBIO-InBIO and Geography Research Unit. The candidate was financially supported by Fundação para a Ciência e Tecnologia (FCT, Portugal) through the attribution of the PhD grant PD/BD/132407. Financial support was given to the project by AGRIGEN-NORTE-01-0145-FEDER- 000007, supported by Norte Portugal Regional Operational Programme (NORTE2020), under the PORTUGAL 2020 Partnership Agreement, through the European Regional Development Fund (ERDF).

To my mother.

Acknowledgements

First, I would like to express my profound gratitude to José Carlos Brito. He was not only a supervisor, in the true meaning of the word, but he impacted my life profoundly. He believed in me when no one else did and had worked hard to allow me to perform always at my best level. He was able to recognize hidden values on me and allowed me to visit and fell Africa as I could never imagined. In the words of Sir Isaac Newton: “If I have seen further it is by standing on the shoulders of Giants”. And what a giant José Carlos Brito is. More than an advisor he is a true friend, and we developed a friendship at levels far from what I could ever had imagined. I am also deeply thankful to Jarkko Saarinen for accepting to co-supervise this Thesis and for teaching me the meaning of many tourism terms. He is highly knowledgeable in definitions and was a precious intervener in all sorts of articles production. The institutions that funded this project deserve a special appreciation: the Rufford Foundation and the Mohammed bin Zayed Conservation Fund (who funded field expeditions), the European Union and the Governments of Portugal and of Finland (who funded Research Centers), the Fundação para a Ciência e Tecnologia and the Academy of Finland (who funded key people for this Thesis). BirdLife International, IUCN, World Bank, United Nations, and many other global institutions must be thanked for providing free data that were key to this Thesis. A special thanks go to the Sahara Conservation Fund. Not only did they provide data for free, but they are also responsible for some of the most beautiful conservation projects out there. They are an inspiration for those wanting to conserve wildlife in the Sahara. I am grateful to the University of Porto and CIBIO-InBIO (in Portugal) and to the University of Oulu and Geography Research Unit (in Finland) for the opportunity to conduct research in these outstanding institutions. I thank the many people within Biodeserts group who allowed me to grow by teaching me about desert ecology and biodiversity and about many other subjects out of the Biology scope. Thanks for the great conversations. In terms of professional assistance, I would like to thank Paulo Pereira. He helped me a lot with his profound knowledge of programming and data analysis, or, when I was feeling desperate with the statistical programs, he helped to burst out laughing. I also thank Pedro Tarroso and Sílvia Carvalho for discussions about specific softwares that I had used along this Thesis. Part of this work was conducted in the field, where I had some of my best moments in life. The desert expeditions were more than amazing because of João Campos, Fábio Sousa, and Joana Marques. Andack Saad Sow helped us a lot in the field and he showed me the ‘Mauritanian way’ – even if I did not understand many of his words, I could understand their meaning. Thank you for showing me your people and your country’ values. Other people that helped me in the field need to be acknowledged as well: Fernando Martínez-Freiría, Guillermo Velo-Antón, Juan M. Pleguezuelos, Olivier Peyre, Phillippe Geniez, Pierre-André Crochet. I thank the many Mauritanians that helped us in a way or another. Professor Dieng Hamidou from the University of Nouakchott, who constantly invite us to perform biological expeditions in ; the Diawling National Park staff that accompanied us in the field; and the many unnamed people that we have met along the way that helped me to better understand Mauritania’ culture and values. Part of this work was conducted in the cold lands of northern Finland (Oulu). Those days were warmed by some people that I met and to whom I am also thankful. Denner Vieira was my best friend there – in just two months we developed such a friendship; thank you so much my friend! Georg from Austria, Roger from USA, and Audrey from Canada: thank you all! Professionally, I also thank Sofia Vaz and João Carvalho for inviting me to great collaborations. These two are the kind of scientists that any institution needs to have among their staff! I am also thankful to many African artists that helped me to write hours on end with their music playing in the background: Rokia Traoré (), Sona Jobarteh (Gambia), Fatoumata Diawara (Mali), Tinariwen (Mali), Bombino (), Tamikrest (Mali), Ballaké Sissoko (Mali), Ali Farka Touré (Mali), Salif Keita (Mali), Noura Mint Seymali (Mauritania), Malouma (Mauritania), Mdou Moctar (Niger), Cheikh Lô (). That this Thesis can help them to lift their countries as they lifted me through their spectacular songs. I thank the many friends that have been accompanying me for years and allowed some of the best laughing moments a PhD student can have. João Moura, Alice Lucas, Diogo Silvério, Miguel Friães, André Gonçalves, Sara Moutinho, Sandra Oliveira, Margarida Henrique, Leonor Gouveia, Pedro Mourão, Gustavo Correia, Sofia Bessa, Ana Catarina Santos, João Manuel, Catarina Santos, and Claúdia. Please apologize if I forgot someone. What about unlikely friends? António Ferreira, Renato Fernandes, Bárbara Barroso, Pedro Silva-Santos, Francisco Teixeira, and Ricardo Teixeira (and all the KIAI family), who have lately added tremendous value through masterminds and mentorships. I thank my family for always being there, no matter what. I am the first to complete the highest level of studies among the closest members and I hope this Thesis can inspire others to follow these steps. Education always wins! And because the best always come at the end, I am profoundly thankful to my mother. She instilled me principles of the greatest value. She gave me the tools for me to thrive. She always believed I can achieve the next level. Not only the next level, but the best level! If I would be allowed only one word of appreciation, I would say my mother was the only responsible for me to be where I am today. She thinks I taught her a lot about Biology. But she taught so much more about life. THANK YOU, MOM! i FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Resumo

Por todo o mundo, os ecossistemas estão ameaçados pelo desenvolvimento humano insustentável. Os desertos, em particular, enfrentam uma das maiores taxas de perda de espécies e de . Contudo, continuam a ser negligenciados pela sociedade global. E apesar de providenciarem uma quantidade considerável de serviços de ecossistemas que beneficiam as pessoas, estão entre os ecossistemas menos estudados e financiados para conservação. Alternativas sustentáveis são urgentemente necessárias para conservar alguns dos últimos biomas selvagens e fornecer às comunidades locais os meios que elas precisam para prosperar no longo termo. O ecoturismo (um dos principais serviços de ecossistema culturais para a Humanidade) tem sido defendido como uma solução potencial para a conservação dos desertos, já que é capaz de preservar espécies carismáticas e não carismáticas enquanto melhora as condições socioeconómicas locais. Como um dos setores do turismo que mais cresce a nível global, o ecoturismo pode ser a resposta para o desenvolvimento local sustentável. Inclusive, tem sido reconhecido pelas Nações Unidas como crucial para a diminuição da pobreza e preservação de vida selvagem em regiões remotas do planeta. Contudo, o seu desenvolvimento em ecossistemas desérticos continua pouco estudado. De entre as maiores regiões áridas do mundo, o Sahara-Sahel é o deserto mais negligenciado. Apesar de ser rico em património natural e cultural que sustenta o desenvolvimento de atividades de ecoturismo, os conflitos regionais e atividades insustentáveis em áreas malgovernadas têm diminuído a capacidade dos países do Sahara e do Sahel se desenvolverem apropriadamente. O ecoturismo poderá ajudar a resolver os conflitos locais enquanto melhora as condições socioeconómicas dos locais e beneficia a conservação de espécies ameaçadas. Porém, a investigação no Sahara- Sahel está ainda a dar os seus primeiros passos. Nesta Tese, estruturei e realizei uma das primeiras avaliações para o desenvolvimento sustentável do Sahara-Sahel. O principal objetivo desta Tese é compreender o papel de espécies-bandeira para a conservação e para o ecoturismo e como é que o ecoturismo pode ser uma solução sustentável para a conservação dos desertos e da melhoria dos modos de vida locais. Especificamente procuro: 1) perceber que espécies podem ser usadas como bandeira para campanhas de conservação e de marketing e que áreas concentram maior número destas espécies-bandeira, 2) entender quais são as áreas mais adequadas para o desenvolvimento do ecoturismo nos principais pontos de água na Mauritânia (escala local) e nos 18 países do Sahara-Sahel (escala subcontinental), 3) avaliar o desempenho dos países em preservar e promover os serviços de ii FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

ecossistema culturais, e 4) compreender os impactos das alterações climáticas nos principais pontos de interesse turístico no Sahara-Sahel. Para poder dar resposta a estes objetivos de investigação, primeiro revi a literatura ligada à investigação em ecoturismo nos desertos, identifiquei o património natural e cultural dos desertos que pode ser usado para promover o ecoturismo, os constrangimentos ao desenvolvimento de atividades de ecoturismo, e apresentei os principais impactos das atividades ecoturísticas nas componentes ecológicas, económicas e socioculturais. Isto originou a introdução da Tese. Para o primeiro objetivo, desenvolvi um novo método para identificar espécies-bandeira em regiões remotas onde o conhecimento é escasso, usando métodos de ordenação e de agrupamento que até à data não tinham sido aplicados na literatura. Esta metodologia baseada em duas fases (ordenação e agrupamento) permitiu avaliar os traços que tornam as espécies atrativas de forma independente e agrupar espécies que partilham características semelhantes em grupos de espécies-bandeira. Esta metodologia permitiu mapear as áreas que concentram elevada riqueza de espécies- bandeira desérticas pela primeira vez, evidenciando o seu potencial para a promoção de campanhas de ecoturismo e de conservação. Para cumprir o segundo objetivo, propus uma nova abordagem que combina procedimentos estatísticos independentes para avaliar o potencial ecoturístico de uma região. Combinar múltiplos critérios com algoritmos de ordenação e de agrupamento permitiu identificar pontos de água na Mauritânia que são adequados para o desenvolvimento de ecoturismo e avaliar de forma independente quais as características que estão mais relacionadas com o potencial ecoturístico. Esta nova abordagem permitiu agrupar locais para diferentes mercados de ecoturismo (ecoturistas “leves” e “duros”), o que permite otimizar investimentos na região analisada. Depois mapeei e analisei os serviços de ecossistema culturais dos vários países do Sahara- Sahel e identifiquei os pontos onde esses serviços culturais beneficiam mais a Humanidade. Esta análise mostrou que os desertos providenciam muito mais serviços de ecossistema do que previamente se pensava e que os países do Sahara-Sahel precisam de desenvolver planos estratégicos transfronteiriços para travar a destruição da natureza e melhorar os modos de vida locais através da promoção do potencial dos ecossistemas locais. Para responder ao terceiro objetivo, usei as mais refinadas e atualizadas ferramentas de planeamento da conservação e dados socioeconómicos atualizados para compreender o desempenho dos vários países do Sahara-Sahel em fornecer e gerir serviços de ecossistema culturais. Isto permitiu identificar prioridades nacionais para a gestão dos ecossistemas e perceber que países estão a perder oportunidades para o FCUP iii Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel desenvolvimento local. A componente multidisciplinar deste trabalhado apresentou uma ferramenta construtiva com que os governos locais podem desenvolver medidas para a gestão eficiente dos ecossistemas. Para cumprir o quarto objetivo, usei dados climáticos e modelos globais atualizados para avaliar os locais mais interessantes para o turismo que poderão ser impactados pelas alterações climáticas. Esta análise permitiu compreender que regiões turísticas estão ameaçadas de perigo perante diferentes cenários de alterações climáticas e que áreas serão insustentáveis para a sobrevivência humana nos desertos. Esta Tese mostrou o potencial do ecoturismo em conservar o património desértico e as oportunidades que os desertos providenciam para desenvolver atividades económicas sustentáveis que beneficiam as comunidades locais. O planeamento e ferramentas geoestatísticas inovadoras usadas nesta Tese permitiram investigar cenários para a sustentabilidade em ambientes desérticos. A componente interdisciplinar desta Tese e os métodos desenvolvidos a múltiplas escalas providenciaram uma ferramenta que a comunidade internacional pode usar para alcançar o desenvolvimento sustentável nos ecossistemas mais negligenciados a nível global. Apesar de que o ecoturismo não pode ser encarado como uma panaceia para todos os desafios que os desertos enfrentam, pode ser visto como uma ferramenta sustentável que contribui positivamente para o bem-estar das comunidades que vivem nesses desertos. Por último, o ecoturismo em desertos pode contribuir para a concretização da Agenda 2030 das Nações Unidas sobre os Objetivos para o Desenvolvimento Sustentável e para a conservação e gestão do património desértico. Esta Tese deu voz a essa reivindicação.

Palavras-Chave

Animais carismáticos, Análises socioeconómicas, Atrações e constrangimentos no deserto, Desenvolvimento sustentável, Desertos, Impactos, Divulgação da conservação, Priorização espacial, Segmentos de ecoturistas, Turismo sustentável iv FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Abstract

Global ecosystems are threatened by human unsustainable development. In particular, deserts are facing some of the highest species and loss rates. Yet, they remain largely neglected in public discussions and are among the least studied and underfunded ecosystems for conservation, despite providing multiple ecosystem services that benefit people. Therefore, alternative and sustainable land-uses are urgently needed for conserving these last wild biomes and provide local people with the needs to prosper in the long term. Ecotourism (one of the key cultural ecosystem services) has been advocated as a potential solution for the desert conservation crisis, able to conserve charismatic and non-charismatic species while improving local socioeconomic conditions. As one of the fastest growing sectors within the tourism field, ecotourism may serve as an answer for sustainable local development. The United Nations also recognise ecotourism as a key for poverty alleviation and wildlife preservation in remote regions, yet its application to desert ecosystems remains little studied. Among the Earth’s largest arid regions, the Sahara-Sahel is the most neglected desert. Despite being rich in natural and cultural heritage that could substantiate the development of ecotourism activities, regional conflicts and unsustainable activities in poorly governed societies and regions have been detracted Sahara and Sahel countries from proper development. Ecotourism might help ameliorate conflicts while improving socioeconomic conditions for local people and benefit the conservation of imperilled species, but its research and development on Sahara-Sahel is still in its early stages. In this Thesis, I performed one of the first ecotourism assessments and frameworks for the sustainable development in Sahara-Sahel. The main goal of the Thesis is to understand the role of flagship species for conservation and ecotourism and how ecotourism could work as a sustainable solution for desert conservation and local livelihoods improvement. In particular, I want to 1) understand which species can be used as flagship for conservation and ecotourism marketing and where are located hotspots of these groups of flagships, 2) assess which are the most suitable areas for ecotourism development in Mauritanian inland water-bodies (local scale) and in the 18 Sahara-Sahel range countries (sub-continental scale), 3) evaluate countries’ performance in preserving and promoting the cultural services that their ecosystems supply, and 4) understand the implications of climate change on regional tourism. In order to answer these research goals, I first overview the literature to present the ecotourism research on deserts, identify the key natural and cultural desert heritage that can be used to promote ecotourism but also the constraints to its development, and the FCUP v Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel impacts of its activities on the ecological, economic and socio-cultural facets. This gave the moat to the introduction of the Thesis. For the first goal, I developed a new method to identify flagship species in remote regions with knowledge gaps, by using statistical approaches (ordination and clustering methods) not used to date in the flagship literature. This two-stage statistical methodology allowed to assess species’ appealing traits in an independent, non-ad-hoc way, and to group species sharing similar characteristics onto flagship fleets. It also allowed to map areas concentrating the highest richness of desert flagships for the first time, highlighting their potential for ecotourism and conservation promotion. To achieve the second goal, I proposed a novel approach that combines independent statistical procedures to assess ecotourism potential. Combining multi-criteria with ordination and clustering algorithms allowed to identify suitable water-bodies for ecotourism development in Mauritania and to independently assess which features are related with ecotourism potential. This new approach enabled to group sites for different ecotourism markets (hard and soft ecotourists), which ultimately allows to optimize investments along the analysed region. I then mapped and analysed cultural ecosystem services along the Sahara-Sahel countries and identified hotpots where cultural services can benefit the Humanity the most. This analysis showed that deserts supply much more ecosystem services than previously thought and that Sahara-Sahel countries need to project transboundary strategic plans, which would halt natural destruction and improve local livelihoods through the promotion of the potential of local ecosystems. For answering the third goal, I used sophisticated and most updated decision-support tools for conservation planning and socioeconomic data to understand the performance of Sahara-Sahel countries in supplying and managing cultural ecosystem services. This allowed to identify national priorities for cultural services management and to identify countries missing opportunities for local development. The interdisciplinary component of this work presented a constructive framework for local governments to shape regional policies towards sustainable ecosystem management. To achieve the fourth goal, I used historical and future climatic data to evaluate which tourism hotspots will be most impacted by climate change. This analysis allowed to understand which touristic regions are in danger from global warming and which areas will be unreliable for human survivability in deserts. This Thesis showed the potential of ecotourism for desert heritage conservation and the opportunities deserts provide to develop sustainable economic activities that benefit local communities. The innovative site-selection planning and geostatistical tools used in this Thesis allowed to thoroughly investigate scenarios for sustainability in threatened desert environments. The interdisciplinary component of this Thesis and the methods vi FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

developed at multiple scales provided a framework that the international community can use for achieving sustainable development in the most neglected global ecosystems. Although ecotourism should not be envisaged as a panacea for all the challenging impacts that deserts are facing, it can offer a sustainable tool to contribute positively to the wellbeing of communities living within deserts. Ultimately, desert ecotourism can contribute to the 2030 agenda of the United Nations for the Sustainable Development Goals and the conservation and management of desert heritage, and this Thesis aimed to give voice to this claim.

Keywords

Charismatic , Conservation marketing, Impacts, Economic analyses, Ecotourism potential, Deserts, Desert attractions and constraints, Ecotourist segments, Spatial prioritization, Sustainable development, Sustainable tourism 1 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table of Contents

Abstract ...... iv Chapter 1 ...... 18 1. General Introduction ...... 19 Article I: Desert Conservation and Management: Ecotourism ...... 31 Chapter 2 ...... 64 2.1. General objectives ...... 65 2.2. Thesis structure ...... 65 Chapter 3 ...... 68 Article II: New method to identify and map flagship fleets for promoting conservation and ecotourism ...... 69 Chapter 4 ...... 105 Article III: Using multivariate statistics to assess ecotourism potential of waterbodies: A case-study in Mauritania...... 106 Article IV: Mapping and analysing cultural ecosystem services in conflict areas ...... 133 Chapter 5 ...... 176 Article V: Assessment and prioritization of cultural ecosystem services in the Sahara- Sahelian region ...... 177 Article VI: Vulnerability of desert tourism to climate change: lessons from the Sahara- Sahel ...... 222 Chapter 6 ...... 256 6. General Discussion ...... 257 Appendices ...... 286

2 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

List of Tables

Table 1.1. Potential attraction and constraining elements of the desert biome to develop ecotourism...... 35 Table 1.2. Potentially positive and negative impacts from ecotourism in desert biome...... 50 Table 2.1. Variables used to evaluate the 1126 Sahara-Sahel species to be used in flagship marketing campaigns, their code, description, type (num: numerical; cat: categorical), units, and data sources...... 77 Table 2.2. Correlation coefficients of species traits with the first two axes (Dimension 1 and Dimension 2; highly correlated values are in bold) of the Principal Component Analysis with mixed data (PCAmix), the eigenvalues and the percentage of explained variation of these two axes. See Fig. 2.1 for the squared loadings of the species traits within the two first rotated axes...... 81 Table 2.3. List of species within flagship fleets suitable for conservation and ecotourism flagship marketing campaigns in the Sahara-Sahel and indication of the class, taxa, common name, endemic status (END; Sahara or Sahel), IUCN conservation status (CS), and flagship hotspot (numbered by location) where they can be observed (see Fig. A1 for details on the areas considered). Species that are distributed outside the flagship hotspots are signalled (Outside). Species data follow IUCN (2017)...... 83 Table 3.1. Types of variables used to assess the ecotourism potential of water-bodies, their codes, description, interest for ecotourism (Min-minimum and Max-maximum), units, and source...... 113 Table 3.2. Loadings of the principal components, and the standard deviation (SD) and cumulative variance (CUM) of the principal components...... 119 Table 3.3. Categories and subcategories of the variables assessed to map and analyse cultural ecosystem services in the Sahara-Sahel, their code, original data types and data sources. * See Table C1 for details...... 141 Table 4.1. Categories of the variables used in the statistical analyses, their codes, and data sources. See Text S1 for a summary description of the variables. The reader is referred to data sources for a complete description and methodological details behind the data here described...... 182 Table 4.2. Loadings of the principal components (PC1-PC17), the variability (VAR) and the cumulative variance (CUM) of the principal components to evaluate the potential performance of countries in managing CES supply. Predictors highly correlated with the two first principal components (that cumulative explain most of the variance) are marked in bold. See Table 4.1 for socio-economic variable codes...... 186 FCUP 3 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table 4.3. Global Climate Models (GCMs) for two different Shared Socio-economic Pathways (SSPs 126 and 585; Riahi et al., 2017) for the years 2081-2100 used in this study and the institutions responsible for developing those GCMs. The codes, model full names (when available), institutions responsible for the models and the main reference are provided. Models and institutional data retrieved from the Coupled Model Intercomparison Project 6 (CMIP6; Eyring et al., 2016) of the World Climate Research Programme (WCRP)’s Working Group of Coupled Modelling (WGCM) at https://www.wcrp-climate.org/wgcm-cmip. For full data sources consult the World Data Center for Climate (WDCC) at https://cera.-www.dkrz.de. Bioclimatic variables for all GCMs were downloaded at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org...... 230 Table A1. The nine flagship fleets identified by the Bayesian Information Criterion (BIC). Description, class and species (Taxa) – alongside their endemism (END; Sahara or Sahel) and IUCN conservation (CS) status – that belong to each of the fleets. Species data follow IUCN (2017)...... 300 Table B1. Main characteristics of the three groups identified by the PCA and their suitability for different types of ecotourists See Figures 2 and 3 for grouping details and geographic distribution of each group, respectively...... 339 Table C1. Types and definitions of the variables chosen to map cultural ecosystem services in the study area...... 344 Table C2. List of flagship species suitable for conservation and ecotourism flagship marketing campaigns in the Sahara-Sahel. Indication of the class, taxa, common name, endemic status (END; Sahara or Sahel), and IUCN conservation status (CS). Species data follow IUCN (2019). Adapted from Santarém et al. (2019)...... 346 Table D1. Categories and subcategories of the variables used in the prioritization rankings, and their original data types. See references for data details...... 365 Table D2. List of the 10 scenarios used in the ranking process. Considerations of costs (weighted at 1x, 100x and 1000x times; see Supplementary Fig. D7), administrative units, administrative priorities, global weights, and normalization of country area are specified. See Moilanen et al. (2014) for complete methodological details of each scenario...... 366 Table E1. Number of tourism hotspots in each country that is (Present) or will be impacted (Shared Socio-economic Pathways: SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017) by dangerous temperature (>30ºC in BIO 01; >45ºC in BIO 05) and will see dramatic precipitation shifts (>1000mm in BIO 12). Bioclimatic variables are coded as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. 4 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Bioclimatic variables were obtained at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org. See country codes in Fig. 5.1. 384

FCUP 5 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

List of Figures

Fig. 1.1. The seven principles of ecotourism (adapted from TIES, 2015)...... 20 Fig. 1.2. Examples of landscape features typical of the desert biome with potential for ecotourism. (A and B) Sossusvlei in Namib-Naukluft National Park (Namibia); (C) Es Sba (Mauritania); (D) Guelb el Makhsar (Mauritania); (E) Tassili n'Ajjer (); (F, G and H) Tadrart (Algeria). Courtesy of José Carlos Brito and Jarkko Saarinen...... 38 Fig. 1.3. Examples of biodiversity features in the desert biome with potential for ecotourism: (A) African golden wolf (Canis anthus); (B) Felou gundi (Felovia vae) and Rock agama (Agama boulengeri); (C) Desert elephant (Loxodonta africana); (D) Namibian giraffe (Giraffa camelopardalis angolensis); (E) Greater hoopoe-lark (Alaemon alaudipes); (F) Brown-necked Raven (Corvus ruficollis); (G) Lanner falcon (Falco biarmicus); and (H) Desert monitor (Varanus griseus). Courtesy of Frederico Santarém and Jarkko Saarinen...... 40 Fig. 1.4. Examples of historical features associated with human occupation of the desert biome with potential for ecotourism: (A) rock-art representation of giraffe (Tadrart; Algeria); (B) rock-art representation of crocodile (Messak Settafet; ); (C) World Heritage Site of Ghadames oasis (Libya), (D) and (E) (Mauritania), ancient stop-points for caravans crossing the Sahara; (F) historical library in (Mauritania)...... 43 Fig. 1.5. Examples of cultural features associated with human occupation of the desert biome with potential for ecotourism: (A) settlement and food-storage systems of OvaHimba people in Opuwo (Namibia); (B) settlement in Ounianga Serir (); (C) oasis (Ziz valley, ); D - riverbank village (Tassili n'Ajjer, Algeria); (E) desert village (Agadez, Niger); (F) trading route (Aïr mountains, Niger); (G) Watering livestock from rock-pool; and (H) excavated well from dry riverbed (Tagant, Mauritania)...... 46 Fig. 1.6. An example of successful cooperation between local communities, research institutions, national park staff and governmental institutions in the promotion of environmental responsible acts in deserts and in early adoption of ecotourism as a means to preserve the West-African crocodile (Crocodylus suchus) in Diawling National Park, Mauritania. Courtesy of Joana Marques...... 57 Fig. 2.1. Squared loadings of species traits within the first and second axes (Dim 1 and Dim 2) of the rotated PCA with mixed data. The PCA is performed based on six numerical variables and seven categorical variables. See Table 2 for the correlation coefficients of the variables within the two rotated axes. 80 Fig. 2.2. Flagship fleets identified by the Bayesian Information Criterion (BIC) applied to the two first rotated axes (Dim 1 and Dim 2) of Principal Component Analysis with mixed 6 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

data (PCAmix; see Text S2 for methodological details). Large-bodied flagships (brown points) and small-bodied flagships (dark blue points) were used to map flagship hotspots...... 82 Fig. 2.3. Flagship Hotspots of large-bodied flagship fleets (above), small-bodied flagship fleets (centre), and both fleets combined (bottom) in the Sahara-Sahel. Areas with higher flagship richness are depicted in orange (above and centre) and dark red (bottom). The extent of Flagship Hotspots covered by current protected areas network (UNEP-WCMC and IUCN, 2018) is indicated by the Protected Areas polygons in the combined map. See Fig. S1 for the names of the areas...... 90 Fig. 3.1. Location of the 30 water-bodies studied (and code number), main settlements with supporting infrastructures (and their names), paved roads, rivers, and mountains. The location of the capital (Nouakchott) is depicted in the small inset. 112 Fig. 3.2. Water-bodies grouped according to the three first components of the Principal Component Analysis (PCA) and Bayesian Information Criterion (BIC). The seven subgroups (blue triangles; dark green, green and light green circles; dark red, red and orange squares) identified by the BIC clustering are within the three groups (blue, green and red hulls) identified by the PCA...... 120 Fig. 3.3. Distribution of the three groups identified by the Principal Component Analysis (PCA; squares, circles and triangles) and of the seven subgroups identified by the Bayesian Information Criterion (BIC) models (different colours of squares, circles and triangles). See Table S1 for grouping details and Figure B2 for grouping examples. 121 Fig. 3.4. Flowchart detailing the steps taken to map and analyse cultural ecosystem services supply in the Sahara-Sahel...... 140 Fig. 3.5. Map of the Sahara-Sahel (a) and distribution of the biodiversity (b-e) and conservation features (f) in the region. BF – ; CAM – ; CAR – ; ER – ; ETH – ; NGA – ; SEN – Senegal; SSD – South ...... 144 Fig. 3.6. Distribution of major landscape features (a-c), caves and gueltas (d), gorges and mountain passes (e) and wetlands and gueltas (f) in the Sahara-Sahel...... 146 Fig. 3.7. Distribution of cultural (a-b) and historical features (c-e) in the Sahara-Sahel...... 148 Fig. 3.8. Distribution of conflict (a-c) and exploitation of natural resources features (d) in the Sahara-Sahel...... 149 Fig. 3.9. Distribution of standardised attractions to cultural ecosystem services supply in the Sahara-Sahel at 0.5-degree resolution. See Table 3.3 for codes of variable...... 152 Fig. 3.10. Distribution of standardised constraints to cultural ecosystem services supply in the Sahara-Sahel at 0.5-degree resolution. See Table 3.3 for codes of variable. .. 153 FCUP 7 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 3.11. Distribution of hotspots of attraction and constraint features to cultural ecosystem services supply in the Sahara-Sahel at 0.5-degree resolution (top and middle) and of hotspots combining many attractions with few constraints and simultaneously many attractions and constraints (bottom)...... 155 Fig. 4.1. Frequency of selection of pixels ranked as low (a) and high (b) priorities for CES conservation on the 10 different scenarios, and the absolute number of planning units (PUs(N)) and the planning units weighted by the area of country within the Sahara-Sahel limits (Dinerstein et al., 2017) (PUs (N/country)) selected for CES conservation (c). Pixels (0.5º resolution) are colored according to the number of times they were selected as priorities in the 10 scenarios: less than two times (grey), between two and eight times (yellow), and more than eight times (blue or red). See Fig. D4 and Table 4.2 for scenarios details, and Fig. D1 for country codes...... 187 Fig. 4.2. Maximum Likelihood Classification of the two first principal components of the spatial PCA (a), and the absolute number of planning units (PUs(N)) and weighted by the area of country within the Sahara-Sahel limits (PUs (N/country)) highly selected for CES conservation (b). Sahara-Sahel ecoregions (Dinerstein et al., 2017) within Africa are depicted in the small inset. Pixels (0.5º resolution) are colored according to the likelihood of belonging to a given class of CES priority raking: blue – low; yellow – middle; red – high. See Figs. 4.1 and D5 for the frequency of selection of low and high priorities, and for the spatial PCA, respectively, and Fig. D1 for country codes...... 193 Fig. 4.3. Results of the Principal Component Analysis to evaluate the potential performance of countries in managing CES. The components that explain most of the variance (PC1 and PC2) are depicted. Countries are represented in blue (see Fig. D1 for country codes) and variables are represented in red (see Table 4.3 for variables codes)...... 194 Fig. 4.4. Relationship between the conditions for visitation and the availability of CES in each country. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socioeconomic indicators are constraining national conditions for CES (Table 4.4 and Fig. 4.3). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country)). Country symbols are colored according to the country-area highly ranked for CES conservation and the 2017 tourism inbounds (UNWTO, 2019): yellow (tourism over-explored); red (tourism under-explored); and black (tourism regularly explored). Symbol shapes represent countries-rankings according to the 2019 index of state of peace (Institute for Economics & Peace, 2019): square – high; circles – medium; diamonds – low; and triangles – very low. Symbol sizes are 8 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

proportional to the raw number of planning units (PUs(N)) highly ranked for CES (see Fig. D6). See Fig. D1 for country codes...... 196 Fig. 4.5. Sensitivity analysis of the relationship between the conditions for visitation and the availability of CES in each country. Conditions for visitation were retrieved from the Human Development Index (HDI) (UNDP, 2019). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country). Country symbols are coded according to the country-area (YY axis) highly ranked for CES conservation and the 2017 tourism inbounds (UNWTO, 2019): yellow (tourism over-explored); red (tourism under-explored); and black (tourism regularly explored). Symbol shapes represent countries-rankings according to the 2019 index of state of peace (Institute for Economics & Peace, 2019a): square – high; circles – medium; diamonds – low; and triangles – very low. Symbol sizes represent the raw number of planning units – PUs(N) – highly ranked for CES conservation (see Fig. D7). Thresholds of medium developed countries (0.55) and high developed countries (0.70) are marked with a dashed line. See Fig. D1 for country codes...... 197 Fig. 4.6. Missing opportunities for local development among Sahara-Sahel countries. Missing opportunities are defined according to the number of planning units weighted by the country-area within the Sahara-Sahel ecoregions (Dinerstein et al., 2017) (PUs (N/country)) prioritized for CES conservation and the inbound tourists in each country. Country symbols are colored as: red (tourism under-explored), yellow (tourism explored to the limit), and black (tourism normally explored when compared to the number of PUs identified with high availability of CES). Tourism data (UNWTO, 2019) are in millions. See Fig. D1 for country codes...... 198 Fig. 4.7. Location of the Sahara-Sahel within North Africa and its countries (top), and the present climate conditions in the study area, depicted by selected bioclimatic variables (bottom). BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic variables were obtained at 10-minute resolution from the WorldClim version 2.1 climate data for the years 1970- 2000 (Fick & Hijmans, 2017) at https://www.worldclim.org. Country codes: ALG: Algeria; BUF: Burkina Faso; CAM: Cameroon; CHA: Chad; EGY: ; ERI: Eritrea; ETH: Ethiopia; LIB: Libya; MAL: Mali; Mauritania: MAU; MOR: Morocco; NIG: Niger; NIA: Nigeria; SEN: Senegal; SSU: South Sudan; SUD: Sudan; and TUN: ...... 228 Fig. 4.8. Averaged and maximum differences between present and future projected climate. Future climate was averaged over eight Global Climate Models – GCMs, for two Shared Socioeconomic Pathways - SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017). Maximum differences were calculated as future FCUP 9 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel averages plus the standard deviation for the eight GCMs for two SSP minus present climatic conditions. Bioclimatic variables are coded as: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic variables were obtained at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org...... 233 Fig. 4.9. Differences between present and future temperature and precipitation levels and their potential impacts on tourism hotspots in Sahara-Sahel. For ease of interpretation, only two thresholds for temperature and four thresholds for precipitation are depicted. Temperatures pushing the limits for visitation are depicted in light red. For precipitation, dark blue depicts changes that may significantly change current ecoregions and, thus, demand for tourists specialised in desert heritage; and light red depicts increased aridity levels. Bioclimatic data for the future were obtained by averaging eight global climate models (GCMs; see Table 4.5) for two Shared Socioeconomic Pathways (SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017). Bioclimatic variables are coded as: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic data for present and the future were obtained at 10-minute resolution from the WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org...... 234 Fig. 4.10. Proportion of tourism hotspots in each country that is (Present) or will be impacted (Shared Socioeconomic Pathways: SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017) by high temperature (>30ºC in BIO 01; >45ºC in BIO 05) and will experience dramatic precipitation shifts (>1000mm in BIO 12). For temperature data, light red depicts where climate will impact tourism hotspots the most. Dark blue depicts precipitation levels high enough to alter ecoregions and landscapes, and impact tourist demand for desert attractions. Lesser impacts are depicted in white (temperature and precipitation data) and light blue bars (precipitation only). Bioclimatic variables are coded as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic variables were obtained at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org. See country codes in Fig. 4.7 and the number of tourism hotspots to be impacted under distinct climate levels in Table E1...... 236 Fig. 4.11. Proportion of tourism hotspots that are currently (Present) or will be (SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017) subject to mean annual temperatures above 30ºC (BIO 01) and maximum temperatures of the hottest month above 45ºC (BIO 05), threatening their visitation. Bioclimatic variables 10 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

were obtained at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org. See country codes in Fig. 4.7...... 237 Fig. 5.1. Examples of flagship species that, due to their morphological and behavioural traits, can attract donors and tourists able to improve local socio-economic conditions in Sahara-Sahel. Photo Credits (clockwise): West-African Crocodile: Frederico Santarém; Addax: Francisco Álvares; Moroccan Spiny-tailed : Zbyszek Boratyński; and Dorcas Gazelle: Zbyszek Boratyński, ...... 259 Fig. 5.2. The West-African crocodile (Crocodylus suchus), a flagship species inhabiting many water-bodies of southern Mauritania that can be marketed in ecotourism campaigns. The picture was taken in 2016 in one of the lagoons with the largest crocodile population in Mauritania. Nearby, there is a village where local people can benefit socially and economically from ecotourism activities. Photo credits: Frederico Santarém...... 261 Fig. 5.3. Local training of Mauritanian people for conservation and ecotourism. Clockwise: JC Brito, 07/11/2010, Mauritania, Sélibabi; JC Campos, 31/10/2011, Mauritania, Tarf Tazazmout; JC Campos, 08/11/2014, Mauritania, Diawling with Zeine of PND; J Marques, 21/08/2015, Mauritania, with team of PN Diawling ...... 268 Fig. A1. Sahara-Sahel limits. Sahara-Sahel limits in Africa (small inset) following Dinerstein et al. (2017), range countries within the area, and main mountains and waterbodies (numbered clockwise). BF – Burkina Faso; CAM – Cameroon; ER – Eritrea; TU - Tunisia...... 298 Fig. A2. Hypothetical representation of ’ body patterns considered in this study...... 299 Fig. B1. Results of the ten Bayesian Information Criterion (BIC) models analysed to identify the most likely number of clusters. Relationships between BIC and number of components are depicted for multiple multivariate models. The model with highest BIC is the spherical, equal volume with 7 components, “EII”. The BIC increases with the number of clusters and most of the increase occurs with less than seven clusters, which were chosen as the optimum number of clusters defining the groups for ecotourism development. Higher number of clusters brings additional complexity to the analysis without a substantial increment of BIC value (Fraley, Raftery, Murphy, & Scrucca, 2012; Fraley & Raftery, 2002)...... 337 Fig. B2. Examples of water-bodies in each of the three main groups identified by the Principal Component Analysis. Green group: water-bodies displaying suitable characteristics for soft ecotourists; Blue group: water-bodies suitable for “hard-desert” ecotourists; and Red group: water-bodies less suitable for ecotourism. Note the great surface area in Green water-bodies, the high topography heterogeneity in Blue water- FCUP 11 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel bodies, and the small size of Red water-bodies. See Table B1 for group and water-body characteristics...... 338 Fig. C1. Sahara-Sahel limits in Africa (small inset) following Dinerstein et al. (2017), range countries within the area, and mountains (yellow balloons), waterbodies (blue balloons), and corridors (red balloons) referred in the text. Adapted from Santarém et al., 2019. BF – Burkina Faso; ER – Eritrea; TU – Tunisia; SEN – Senegal...... 343 Fig. D1. Sahara-Sahel limits in Africa (small inset), its range countries (three-letter codes), the main mountains (yellow balloons), and wetlands (blue balloons) are depicted. ALG: Algeria; BUF: Burkina Faso; CAM: Cameroon; CHA: Chad; EGY: Egypt; ERI: Eritrea; ETH: Ethiopia; LIB: Libya; MAL: Mali; Mauritania: MAU; MOR: Morocco; NIG: Niger; NIA: Nigeria; SEN: Senegal; SSU: South Sudan; SUD: Sudan; and TUN: Tunisia. Figure adapted from Santarém et al. (2020)...... 354 Fig. D2. Relationship between the number of inbound tourists (UNWTO, 2019) in Sahara-Sahel countries and their rank in the Human Development Index (HDI) (UNDP, 2019). Tourism data are for 2017 and are represented in millions. No data are available for South Sudan and thus are not represented in the graph...... 355 Fig. D3. Analytical workflow used to evaluate the performance in managing CES among countries within Sahara-Sahel. The framework combines spatially explicit features and costs (point, polygon, and polyline data at 0.5º resolution) used to prioritize the landscape (A) with predictors of performance in preserving cultural ecosystem services among countries within Sahara-Sahel ecoregions (B). The framework produces spatially and statistically explicit outputs from which it can be identified which countries need to improve touristic, economic, governance, environmental, health, security, and socioeconomic conditions to attract additional international tourists and which ones are missing opportunities given the high potential of their ecosystems to supply cultural services. See Tables 1 and 3 for a detailed explanation of categories used in the prioritization rankings and in the statistical analyses, respectively...... 356 Fig. D4. Sensitivity analyses for different cost weightings (1x, 100x, and 1000x) on 10 different ranking scenarios (a-j). Pixels (0.5º resolution) are colored by ranking priorities of CES availability: blue – low; yellow – middle; and red – high priority for CES conservation. See Table 2 for a detailed description of the 10 scenarios...... 357 Fig. D5. Spatial variability of the high prioritized cells for CES conservation in the Sahara- Sahel. The composite map represents the two first components (PC1 – green, PC2 – blue) of the spatial PCA (89% of the variance explained), which are colored according to the availability of CES (PC1 – green; high priority; PC2 – blue; low priority) at 0.5º resolution. See Fig. 4.1 for the frequency of selection of high priority rankings...... 358 12 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D6. Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socioeconomic indicators are constraining national conditions for CES conservation (Fig. 4.3 and Table 4.4). See Fig. D1 for country codes ...... 359 Fig. D7. Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation. Conditions measured by the Human Development Index (HDI) (UNDP, 2019). See Fig. D1 for country codes...... 360 Fig. D8. Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation and the Zonation prioritization ranking. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socio-economic indicators are constraining national conditions for CES (Fig. 4.3 and Table 4.4). See Fig. 4.1 for Zonation rankings. See Fig. D1 for country codes...... 361 Fig. D9. Sensitivity analysis of the relationship between the conditions for visitation and the availability of CES in each country considering the Zonation prioritization ranking. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socio-economic indicators are constraining national conditions for CES (Fig. 4.3 and Table 4.4). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country)) highly ranked for CES conservation. See Fig. D1 for Zonation rankings and Fig. D1 for country codes...... 362 Fig. D 10. Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation and the Zonation prioritization ranking. Conditions were measured by the Human Development Index (UNDP, 2019). Thresholds of medium developed countries (0.55) and high developed countries (0.70) are marked with a dashed line. See Fig. 4.1 for Zonation rankings and Fig. D1 for country codes...... 363 Fig. D 11. Sensitivity analysis of the relationship between the conditions for visitation and the availability of CES in each country considering the Zonation prioritization rankings. Conditions for visitation were retrieved from the Human Development Index (HDI) (UNDP, 2019). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country)) highly ranked for CES conservation. Thresholds of FCUP 13 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel medium developed countries (0.55) and high developed countries (0.70) are marked with a dashed line. See Fig. D1 for country codes...... 364 Fig. E1. Averages of the eight global climate models (GCMs; see Table 5.1) for two different Shared Socio-economic Pathways (SSP 126 and SSP 585; Riahi et al., 2017) projected for the future for the three bioclimatic variables used in this study. Bioclimatic data are as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Data for the future were obtained at 10-minute resolution from the WorldClim version 2.1 climate data for the years 2081-2100 (Fick & Hijmans, 2017) at https://www.worldclim.org...... 382 Fig. E2. Uncertainty (standard deviation) in future averages of the eight global climate models (GCMs; see Table 5.1) for two different Shared Socio-economic Pathways (SSP 126 and SSP 585; Riahi et al., 2017) projected for the future for the three bioclimatic variables used in this study. Bioclimatic data are as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Data for the future were obtained at 10-minute resolution from the WorldClim version 2.1 climate data for the years 2081-2100 (Fick & Hijmans, 2017) at https://www.worldclim.org...... 383

14 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Abbreviations and Acronyms

AA Area covered by agricultural land ACTIV Activity ALG Algeria AOC Area of occupancy AREA Surface Area Attac Attacks/battles and violence against civilians (2011-2016) BEHAV Behaviour BI Bird species richness index BIO Biocapacity BSIZE Body size BWEIG Body weight BUF Burkina Faso BUS Ease of doing business index CAM Cameroon CarRo Caravan routes CarVi Caravan villages Caves Caves CHA Chad CES Cultural ecosystem services COLPAT Colour pattern CON Underfunded countries (conservation spending) CR Maximum number of crocodiles reported CRM Cultural use CS Conservation DA Area covered by dunes DI Distance to the nearest settlement with supporting infrastructures DR Distance to the nearest paved road Ecore Terrestrial ecoregions ELE Access to electricity (% of population) EGY Egypt END Endemic ERI Eritrea Endem Endemic richness ETH Ethiopia EthnG Major ethnographic groups* ExtraF Oil, gas, mining extractive facilities FEED Feeding Flags Flagship richness Fores Forest reserves Forti Fortifications from colonial period GEF Government effectiveness Global National Index per capita based on purchasing power parity GNI (PPP) GorMt Gorges and mountain passes FCUP 15 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

GPI Global peace index GTI Global terrorism index Guelt Rock pools (known as guelta) Herit UNESCO World Heritage Sites HII Human influence index HIV Prevalence of HIV, total (% of population ages 15-49) INT Individuals using the Internet (% of population) IUCN Number of IUCN threat factors categories LandF Major landscape features* Landm Landmines LIB Libya MAU Mauritania MAL Mali Migra Smuggling and human migration routes Monum Monuments MOR Morogcco MORPH Morphology MPI Global multidimensional poverty index MTWM Maximum temperature of the warmest month NDVI Vegetation Productivity NIA Nigeria NIG Niger NWB Euclidian distance to the nearest water-body NCOL Colour number Oases Oases RockA Rock art Ruins Ruins, tombs, sites of empires historical land occupation Pipel Pipelines ProtA Protected Areas RH Habitat heterogeneity from the nearest paved road ROA Mortality caused by road traffic injury (per 100,000 people) RockF Peculiar rock formations SEAS Seasonality SEN Senegal SL Topographic heterogeneity Speci Total species SSU South Sudan SUD Sudan THR Number of threatened species TIES The International Ecotourism Society TOU International tourism, number of arrivals TUB Incidence of tuberculosis (per 100,000 people) TUN Tunisia VIS Passport Index (countries accepting passports VISA-free) WA Water availability 16 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Mortality rate attributed to unsafe water, unsafe sanitation, and lack of WAS hygiene (per 100,000 population) Wetla Major wetlands WBH Habitat heterogeneity

FCUP 17 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

“If I have seen further it is by standing on the shoulders of Giants” Isaac Newton, 1675

18 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Chapter 1

General Introduction

Article I Santarém, F., Saarinen, J., Brito, J.C. (2019). Desert Conservation and Management: Ecotourism. In: Goldstein MI, DellaSala DA (Eds.), Encyclopedia of the World's Biomes, vol. 2. Elsevier, pp. 259-273

FCUP 19 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

1. General Introduction

1.1. History and definition of ecotourism Tourism and nature conservation have a long symbiotic history. Yellowstone became the first national park in the world to protect nature and for all people to enjoy the unique geologic heritage. But as tourism has grown in natural areas, increasing challenges for sustainability arose. Ecotourism appeared as a refinement of the connection between tourism and conservation (Gössling, 1999). Because of its promise to achieve both conservation and economic development, ecotourism emerged as a form of alternative tourism that emphasized the well-being of host communities in natural areas where the environment needed to be preserved (Stronza, 2019). Yet, the origin of the term ecotourism has been long debated among scholars. The term ‘ecotour’ was originally coined by Canadian Parks in the 1960’s (Fennell, 1999). It appeared later in the English literature as a hyphenated term (eco-tourism) in an article published in the International Journal of Environmental Studies (Romeril, 1985). In the 1990’s the term started to become popularized first by the Mexican ecologist Hector Ceballos-Lascuraín (the original term was in Spannish - ecoturismo; Ceballos-Luscarain, 1992). But it took until the early 2000’s for the term to emerge as we know it today. An official international definition was adopted in 2002 during the United Nations International Year of Ecotourism (UNEP and WTO, 2002), a key mark for the boom of ecotourism research that we had assisted since then. In that same year, the Journal of Ecotourism was established, attracting dozens of new articles focused on the emerging field of ecotourism studies. Since then, several definitions of ecotourism have been revised in over 30 publications dedicated solely to the subject (e.g. Boo, 1990; Blamey, 1997; Buckley, 1994; Cater and Lowman, 1999; Donohoe & Needham, 2006; Diamantis, 1999; Fennel, 2001a). Fast forward to today, and the most accepted and widely used definition is the one provided by The International Ecotourism Society (TIES): “Ecotourism is the responsible travel to natural areas that conserves the environment, sustains the well-being of the local people, and involves interpretation and education” (TIES, 2015). Yet, the principles of ecotourism are reliant among different definitions (Fig. 1.1). 20 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 1.1. The seven principles of ecotourism (adapted from TIES, 2015).

These natural, sociocultural, learning, and sustainable components are extensively discussed in key books about ecotourism (e.g. Buckley, 2009a; Weaver, 2001a, b; and Fennell, 2020).

1.2. Differences to other forms of tourism Ecotourism is often conflated with other types of tourism (Stronza et al., 2019). Terms such as outdoor recreation, nature-based tourism, wildlife tourism, cultural tourism, adventure tourism and even farm tourism have been erroneously used interchangeably with ecotourism (Buckley, 2009a; Weaver, 2001a), especially among biologists and ecologists. Greenwashing of some companies (i.e. by labelling their activities as ecotours when they are not; Honey, 2008) or misunderstanding of the differences between the different types of tourism had led to many unjustifiable critics to ecotourism by academics (e.g. Kiss, 2004). By falling to distinguish ecotourism from other forms of tourism, scholars risk dismissing the conservation purpose of ecotourism (Stronza et al., 2019). The critical issue that distinguishes ecotourism from mostly all other types of tourism is the accepted potential positive contribution to the conservation of the natural environment (Buckley, 2009a; Fennell, 2020; Stronza et al., 2019). Still, a further comment needs to be made to distinguish these concepts. Starting with nature-based tourism, the alternative most common confused with ecotourism by academics, we can see why the terms had been generating so much FCUP 21 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel confusion and misunderstanding. Simply put, nature-based tourism is any type of tourism that relies on natural attractions (Fennell, 2020; Weaver, 2001a). It includes activities scenic sightseeing, wildlife tourism (although Newsome et al. (2013) distinguish this form of tourism as “the viewing of, and non-consumptive encounters with, wildlife solely in natural areas”), as well as visits to botanical gardens, aquariums and zoos, and extractive activities such as hunting and fishing (Buckley, 2009a). It does not necessarily have the learning component, the cultural dimension or even the sustainability character of ecotourism. Yet, ecotourism can be regarded as a subset of nature-based tourism (Fennell 1999). More recently, interesting contests over zoos and hunting/fishing as ecotourism products have been made, amplifying the difference between ecotourism and of nature-based only products (Fennell, 2013). Outdoor recreation is often conflated with ecotourism, but the only communality is the experience that results from recreational activities occurring in natural environments (Moore & Driver, 2005). Ecotourism expands this idea to include support for wildlife in protected areas, diversified livelihoods, environmental education and interpretation, and the strengthening of resource management institutions (Stronza et al., 2019). Adventure tourism activities incorporate an element of risk, higher levels of physical exertion, and the need for certain specialized skills to safely participate in those activities (Huddart and Stott, 2020; Weaver, 2001b). Some forms of ecotourism can qualify as adventure tourism, especially if done in steep mountains, remote deserts, or dense forests that require being in good physical condition. Like nature-based tourism, there is no requirement for sustainability (Buckley, 2010). And the participants of adventure tourism activities seek an environment that facilitates risk and is physically challenging, whereas ecotourists seek more the learning experience (Weaver, 2001a). For a more comprehensive revision on these topics, there are several books discussing the main similarities and differences between the sectors. For example: nature-based tourism: Hall & Boyd (2005); wildlife tourism: de Lima & Green (2017); Newsome et al. (2005); adventure tourism: Buckley (2006, 2010), Huddart and Stott (2020); Taylor et al. (2013); Sznadjer et al. (2009); ecotourism: Buckley (2009a); Diamantis (2004); Honey (2008); Weaver (2001a, 2001b).

1.3. Benefits of ecotourism As seen before, ecotourism expands the contributions of its products/activities for sustainable development. It addresses both environmental and social goals, having been promoted by the United Nations as an excellent option to achieve the 2030 22 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Sustainable Development Goals (UN, 2018). I briefly present some of the ecotourism benefits under; a more comprehensive discussion on the ecotourism impacts upon deserts is presented in the following article.

1.3.1. Conservation benefits One of the key documented conservation benefits of ecotourism is the protection of threatened species. Many works have been demonstrating the key role of ecotourism in preserving wild species (e.g. Buckley et al., 2016, 2012; Lindsey, 2005; Mossaz et al., 2015; Nevin & Gilbert, 2005; Steven et al., 2013) and natural places (e.g. Brandt & Buckley, 2018) around the world. For example, ecotourism’ positive impacts at the landscape-level have been documented in ’s Ngorongoro Crater Conservation Area (Charnley, 2005); Peru’s Tambopata National Reserve (Kirkby et al., 2011, 2010); and Ecuador’s Galapagos Islands National Park (Powels & Ham, 2008). Additionally, an indirect way in which ecotourism can work for conservation is by strengthening local institutions (Stronza et al., 2019), particularly environmental NGOs (Romero-Brito, 2016). There is increased evidence that tourism works best for conservation when it is based on the principles and practices of ecotourism (Brandt & Buckley, 2018).

1.3.2. Economic benefits Ecotourism financial benefits extend largely, from the leading of new sources of income to betterment of household conditions and improvement of local livelihoods (Das & Chatterjee, 2015). Job creation in tourism services such as tour guiding services, accommodation, transportation, food provision and handicraft provides an economic mean to the many local communities living in remote places where ecotourism takes place (Stronza, 2007; Reimer & Walter, 2013; Wunder, 2000). For instance, in Osa Peninsula, Costa Rica ecotourism offers the best available employment opportunities (Hunt et al., 2014) and the financial benefits in the rainforest of Tambopata region of Amazonian Peru is higher than any other alternative (Kirkby et al., 2010). Economic benefits extend to the creation of new jobs in sectors not directly related to ecotourism, but that benefit from ecotourism promotion (Hunt et al., 2014). This is the case of the consumer goods sector that benefits from the increased purchasing power of local people employed in ecotourism (Das & Chatterjee, 2015). Ultimately, ecotourism increases levels of financial support for environmental conservation (Cisneros- Montemayor et al., 2013; Hunt et al., 2014) and can improve the quality of services for visitor reception at national parks (Ly et al., 2006). Ecotourism contributes to diversity the livelihoods of people living in places where activities take place (Ferraro & Hanauer, 2014, 2011). Because locals that are dependent FCUP 23 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel on local environments lessen their reliance on natural resources when integrated into ecotourism, conservation benefits can be added to the economic benefits. Examples of these were found in Guatemala (Langholz, 1999), in Ecuador (Wunder, 2000), and in Costa Rica (Troëng & Drews, 2004). But this is not always the case, as found in Nepal, Mexico, and Belize (Bookbinder et al. 1998, Barkin, 2003, and Lindberg et al. 1996, respectively).

1.3.3. Sociocultural benefits Ecotourism empowers local communities socio-politically and fosters respects for different human cultures (Das & Chatterjee, 2015). It has been shown to contribute to cultural pride among host communities, as well as to support local arts, revitalize ethnic traditions, customs, shared identities, and even languages (Coria & Calfucura, 2012; Stronza, 2008; Zeppel, 2006). In particular, community-based ecotourism has been promoted as a sub-field of ecotourism that focuses on the economically, socially, and culturally empowering of local communities through participation (Reimer & Walter, 2013). Ecotourism is also claimed to promote gender equality and to motivate young girls to go back to school (Horton, 2009; Scheyvens, 2000), although the opposite view is also discussed in some studies (Badola & Hussain, 2003; Scheyvens, 2000).

1.4. Ecotourism segmentation for marketing Ecotourists in general are well educated with a tertiary education and a high income level (Buckley, 2009a, Weaver, 2001a,b), which results in a higher willingness to spend money in the destination country (Wight 1996). Different types of ecotourists contribute to these expenditures and market segmentation divides them into “hard” and “soft” ecotourists according to their motivations, attitudes, and behaviors. Hard ecotourists are considered as having strong biocentric attitudes and commitments to environmental issues. They seek deep and meaningful interactions with the environment, travelling in small, specialized groups, and actively search for physically challenging experiences requiring few or no tourism supporting infrastructures. They emphasize more the personal experience and tend to make own travel arrangements. As such, the hard ecotourist may spend more time with local communities to understand their relationships with nature. On the opposite side, soft ecotourists tend to prefer experiences of shorter duration with superficial environmental commitment, in larger groups of general tourism interests. They seek mostly physically passive and comfort activities and expect tourism services. They 24 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

present shallow interactions with nature but have a strong emphasis on interpretation rather than on personal experience. They rely mostly on travel agencies and tours to arrange their activities (Weaver, 2005, 2001a; Weaver & Lawton, 2002). Market segmentation is, therefore, important for ecotourism developers as it helps to manage the natural resources and their utilization. It is also important to understand ecotourists demand for different activities and the sectors that are more active in environmental and conservation initiatives (Weaver & Lawton, 2007).

1.5. Ecotourism research agenda Knowledge in ecotourism is largely dedicated to the environmental, socioeconomic, and cultural impacts of ecotourism on local environments and people, but very few studies had focused on more than one of them at the same time (Stronza et al., 2019). A few other studies were dedicated to the ecotourism supply side of ecosystems and, to a lesser extent, to the demand side (Weaver & Lawton, 2007). Still, there is an overall shortage of sound empirical data (Brandt & Buckley, 2018; Fennel, 2001b) and very few ecotourism articles employ multivariate analyses, necessary to develop a strong knowledge base (Weaver, 2001a). Many works did not access the role of countries in managing their cultural ecosystem services or the external environments that might impact ecotourism supply and demand, such as terrorism, overexploitation of natural resources and conflicts on susceptible regions (Weaver & Lawton, 2007). Incorporating broader socio-ecological and political analyses into ecotourism studies is timely needed (Stronza et al., 2019). Another critical issue in ecotourism analysis is related to scale (Buckley, 2009b). Most studies are performed at small or large scales, but very few combined multiple scales and assessed scale-effects for ecotourism (Stronza et al., 2019; Hall, 2007). Additionally, almost no studies addressed the bifurcation of ecotourism into hard and soft dimensions. This is a prominent field in the ecotourism literature that has implications for the segmentation and management of local heritage where ecotourism activities take place and needs to be further explored (Weaver & Lawton, 2007, 2002). Of a larger importance, is the ecotourism literature specialization in the different world biomes. Ecotourism studies are abundant in tropical forests and savannahs but are lacking to a larger extent in deserts and arid lands (Kruger, 2005; Santarém and Paiva, 2015; Weaver, 2001b). Even when analysed at the level of cultural ecosystem services, research in deserts is far behind other biomes (Lu et al., 2018). To emphasize this knowledge gap, I present here a literature revision of the ecotourism research on deserts (Article I). I review the potentials and constrains of FCUP 25 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel deserts for ecotourism development as well as the impacts ecotourism might derive to the desert ecological, economic, and socio-cultural components. 26 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

References Badola, R., & Hussain, S.A. (2003). Conflict in paradise: Women and protected areas in the Indian Himalayas Mountain. Research and Development, 23(3), 234–237. Barkin D. 2003. Alleviating poverty through ecotourism: promises and reality in the Monarch Butterfly Reserve of Mexico. Environment, Development, and Sustainability, 5(3/4): 371– 382. Blamey, R. (1997). Ecotourism: The search for an operational definition. Journal of Sustainable Tourism, 5, 109–130. Boo, E. (1990). Ecotourism: The Potentials and Pitfalls. Worldwide Fund for Nature: Baltimore, USA. Bookbinder, M.P., Dinerstein, E., Rijal, A., Cauley, H., Rajouria, A. (1998). Ecotourism’s support of biodiversity conservation. Conservation Biology, 12(6): 1399–1404. Brandt, J. S., & Buckley, R. C. (2018). A global systematic review of empirical evidence of ecotourism impacts on forests in biodiversity hotspots. Current Opinion in Environmental Sustainability, 32: 112-118. Buckley, R. (2010). Adventure Tourism Management. Elsevier: Oxford, UK. Buckley, R. (2009a). Ecotourism: Principles & Practices. CABI: Wallingford, UK. Buckley, R. (2009b). Evaluating the net effects of ecotourism on the environment: a framework, first assessment and future research. Journal of Sustainable Tourism, 17(6), 643–672. Buckley, R. (2006). Adventure Tourism, CABI: Wallingford, UK. Buckley, R. (Ed.). (2004b). Environmental impacts of ecotourism. CABI: Wallingford, UK. Buckley, R. (1994). A framework for ecotourism. Annals of Tourism Research, 21: 661-669. Buckley, R., Castley, J. G., Pegas, F. de V., Mossaz, A. C., & Steven, R. (2012). A population accounting approach to assess tourism contributions to conservation of IUCN-redlisted mammal species. PLoS ONE 7: e44134. Buckley, R., Morrison, C., & Castley, J. G. (2016). Net effects of ecotourism on threatened species survival. PLoS ONE 11: e0147988. Cater, E. & Lowman, G. (Eds.) (1994). Ecotourism: A Sustainable Option? John Wiley & Sons: Chichester, UK. Ceballos-Luscaráin, H. (1992). Tourism, ecotourism and protected areas. Proceedings o IV Congress on National Parks and Protected Areas. IV Congress of National Parks and Protected Areas, Caracas. IUCN, Gland, Switzerland. Charnley S. (2005). From nature tourism to ecotourism? The case of the Ngorongoro Conservation Area, Tanzania. Human Organization, 64(1):75–88. FCUP 27 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Cisneros-Montemayor, A. M., Barnes-Mauthe, M., Al-Abdulrazzak, D., Navarro-Holm, E., & Sumaila, U. R. (2013). Global economic value of shark ecotourism: implications for conservation. Oryx, 47(03), 381–388. Coria J, Calfucura E. 2012. Ecotourism and the development of indigenous communities: the good, the bad, and the ugly. Ecological Economics, 73: 47–55. Das, M. Chatterjee, B. (2015). Ecotourism: A panacea or a predicament? Tourism Management Perspectives, 14: 3-16. Diamantis, D. (2004). Ecotourism: Management and Assessment. Thomson Learning: Cornwall, UK. Diamantis, D. (1999). The concept of ecotourism: evolutions and trends. Current Issues in Tourism, 2: 515-518. Donohoe, H. M., & Needham, R. D. (2006). Ecotourism: The Evolving Contemporary Definition. Journal of Ecotourism, 5(3), 192–210. Fennell, D. A. (2020). Ecotourism (5th Edition). Routledge: Abingdon, UK. Fennell, D. A. (2013). Contesting the zoo as a setting for ecotourism, and the design of a first principle. Journal of Ecotourism, 12(1): 1-14. Fennell, D. (2001a). A content analysis of ecotourism definitions. Current Issues in Tourism, 4, 403–421. Fennel, D. (2001b). Areas and Needs in Ecotourism Research. In Weaver (2001). The Encyclopedia of Ecotourism. CABI: Wallingford, UK. Fennell, D. (1999). Ecotourism: An Introduction. Routledge: London, UK and New York, USA Ferraro, P.J., Hanauer, M.M. (2014). Quantifying causal mechanisms to determine how protected areas affect poverty through changes in ecosystem services and infrastructure. PNAS 111(11): 4332–4337. Ferraro, P.J., Hanauer, M.M. (2011). Protecting ecosystems and alleviating poverty with parks and reserves: “win-win” or tradeoffs? Environmental Resources Economics, 48(2): 269–286. Gössling S. (1999). Ecotourism: a means to safeguard biodiversity and ecosystem functions? Ecological Economics 29(2): 303–20. Hall, C. M. (2007). Scaling ecotourism: the role of scale in understanding the impacts of ecotourism. In Higham, J. (Ed.). Critical Issues in Ecotourism: Understanding a complex tourism phenomenon. Routledge: London, 464 pp. Hall, C. M. and Boyd, S. (2005). Nature-based Tourism in Peripheral Areas: Development or Disaster? In Cooper, C., Hall, C. M., and Timothy, D. (Eds.) Aspects of Tourism 21. Channel View Publications: Clevedon, UK. 28 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Honey, M. (2008). Ecotourism and Sustainable Development: Who Owns Paradise? Island Press: Washington D.C., USA. Horton, L.R. (2009). Buying up nature: Economic and social impacts of Costa Rica's ecotourism boom. Latin American Perspectives, 166(36), 93–107. Huddart, D. and Stott, T. (2020). Adventure Tourism: Environmental Impacts and Management. Palgrave Macmillan: Cham, Switzerland. Hunt CA, Durham WH, Driscoll L, Honey M. (2015). Can ecotourism deliver real economic, social, and environmental benefits? A study of the Osa Peninsula, Costa Rica. Journal of Sustainable Tourism, 23(3): 339–57. Kirkby, C.A., Giudice-Granados, R., Day ,B., Turner, K., Velarde-Andrade, L.M. (2010). The market triumph of ecotourism: an economic investigation of the private and social benefits of competing land uses in the Peruvian Amazon. PLoS ONE 5(9): e13015. Kirby, C.A., Giudice, R., Day, B., Turner, K., Silvera Soares-Filho, B., Oliveria-Rodrigues, H., Yu, D.W. (2011). Closing the ecotourism conservation loop in the Peruvian Amazon. Environmental Conservation 38(01): 6–17. Kiss, A. (2004). Is community-based ecotourism a good use of biodiversity conservation funds? Trends in Ecology and Evolution, 19 (5): 232-237. Langholz, J. (1999). Exploring the effects of alternative income opportunities on rainforest use: insights from Guatemala’s Maya Biosphere Reserve. Society & Natutal Resources, 12(2):139–49 Lindberg K, Enriquez J, Sproule K. (1996). Ecotourism questioned: case studies from Belize. Annals of Tourism Research, 23(3): 543–562. Lindsey, P. A., Alexander, R. R., du Toit, J. T., & Mills, M. G. L. (2005). The potential contribution of ecotourism to African wild dog Lycaon pictus conservation in South Africa. Biological Conservation, 123(3), 339–348. de Lima, I. B., and Green, R. J. (Eds.) (2017). Wildlife Tourism, Environmental, Learning and Ethical Encounters: Ecological and Conservation Aspects. Springer: Cham, Switzerland. Lu, N., Wang, M., Ning, B., Yu, D., Fu, B. (2018). Research advances in ecosystem services in drylands under global environmental changes. Current Opinion in Environmental Sustainability, 33: 92–98. Ly, O. K., Bishop, J. T., Moran, D. Dansokho, M. (2006). Estimating the value of ecotourismin the Djoudj National Bird Park in Senegal. Gland, Switzerland: IUCN, 34 pp. Moore, R.L., Driver, B.L. (2005). Introduction to Outdoor Recreation: Providing and Managing Natural Resource Based Opportunities. Venture Publishing: Pennsylvania, USA. FCUP 29 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Morrison C, Simpkins C, Castley JG, Buckley RC (2012) Tourism and the Conservation of Critically Endangered . PLoS ONE, 7(9): e43757. Mossaz, A. C., Buckley, R., & Castley, J. G. (2015). Ecotourism Contributions to Conservation of African Big Cats. Journal for Nature Conservation, 28,:112–118. Nevin, O., & Gilbert, B. (2005). Measuring the cost of risk avoidance in brown bears: Further evidence of positive impacts of ecotourism. Biological Conservation, 123, 453–460. Newsome D, Moore SA, Dowling RK. (2013). Natural Area Tourism: Ecology, Impacts and Management. Channel View Publications: Bristol, UK. Newsome, D., Dowling, R. K., Moore, S. A. (2005). Wildlife Tourism. In Cooper, C., Hall, C. M., Timothy, D. (Eds.) Aspects of Tourism 24. Channel View Publicaitons: Clevedon, UK. Powell R. B., & Ham S. H. (2008). Can ecotourism interpretation really lead to pro- conservation knowl-edge, attitudes and behaviour? Evidence from the Galapagos Islands. Journal of Sustainable Tourism, 16(4):467–89. Reimer, J. K., & Walter, P. (2013). How do you know it when you see it? Community-based ecotourism in the Cardamom Mountains of southwestern Cambodia. Tourism Management, 34, 122–132. Romero-Brito TP, Buckley RC, Byrne J. (2016). NGO partnerships in using ecotourism for conservation: systematic review and meta-analysis. PLOS ONE 11(11):1–19 Santarém, F., & Paiva, F. (2015). Conserving desert biodiversity through ecotourism. Tourism Management Perspectives, 16, 176-178. Scheyvens, R. (2000). Promoting women's empowerment through involvement in ecotourism: Experiences from the third world. Journal of Sustainable Tourism, 8(3), 232– 249. Steven, R., Castley, J. G., & Buckley, R. (2013). Tourism Revenue as a Conservation Tool for Threatened Birds in Protected Areas. PLoS ONE, 8(5), 1–8. Stronza, A. L., Hunt, C. A., & Fitzgerald, L. A. (2019). Ecotourism for Conservation? Annual Review of Environment and Resources, 44(1), 229-253. Stronza, A. (2008). Through a new mirror: reflections on tourism and identity in the Amazon. Human Organization 67(3): 244–57. Stronza, A. (2007). The economic promise of ecotourism for conservation. Journal of Ecotourism, 6(3): 210–221. Sznadjer, M., Przezborska, L., and Scrimgeour, F. (2009). Agritourism. CABI: Wallingford, UK. Taylor, S., Varley, P., Johnston, T. (2013). Adventure Tourism: Meanings, experience and learning. Routledge: Abingdon, UK. 30 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

The International Ecotourism Society (TIES). (2015). What Is Ecotourism? Accessed on 2 September 2020 at https://ecotourism.org/ties-overview/ Troëng, S., Drews, C. (2004). Money talks: economic aspects of marine use and conservation. WWF International. Accessed on 25 August 2020 at http://wwf.panda.org/wwf_news/?153802/wwwpandaorglacmarineturtlespublications. UNEP and WTO (2002). Quebec Declaration of Ecotourism. Accessed on 2 September 2020 at https://www.gdrc.org/uem/eco-tour/quebec-declaration.pdf. United Nations (UN) (2018). Promotion of sustainable tourism, including ecotourism, for poverty eradication and environment protection: Note by the Secretary-General. Accessed on 25 August 2020 at https://digitallibrary.un.org/. Weaver, D. B. (2005). Comprehensive and minimalist dimensions of ecotourism. Annals of Tourism Research, 32(2): 439-455. Weaver, D. B. (2001a). Ecotourism. Wiley: Brisbane, Australia. Weaver, D. B. (2001b). The Encyclopedia of Ecotourism. CABI: Wallingford, UK. Weaver, D. B., & Lawton, L. J. (2007). Twenty years on: The state of contemporary ecotourism research. Tourism Management, 28(5): 1168-1179. Weaver, D. B., & Lawton, L. J. (2002). Overnight ecotourist market segmentation in the Gold Coast hinterland of Australia. Journal of Travel Research, 40: 270–280. Wight, P. (1996). North American ecotourists: market profile and trip characteristics. Journal of Travel Research, 34: 2–10. Wunder, S. (2000). Ecotourism and economic incentives—An empirical approach. Ecological Economics, 32(3): 465–479. Zeppel, H. (2006). Indigenous Ecotourism: Sustainable Development and Management. CABI: Wallingford, UK. FCUP 31 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Article I: Desert Conservation and Management: Ecotourism

Frederico Santarém 1,2,3, Jarkko Saarinen 3,4, José C. Brito 1,2

1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 – Geography Research Unit, University of Oulu, Finland 4 – School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

Abstract

Deserts harbor unique natural and cultural heritage that is found nowhere else in the world. However, threats to these heritage resources are rising due to increasing levels of accessibility, human population, and exploitation of natural resources. Thus, new alternative tools are needed to safeguard these fragile environments. In this respect, ecotourism has the potential to preserve the desert biome while supporting local cultures, traditional livelihoods, and sustainable development. Deserts display many opportunities to promote ecotourism, such as unique biodiversity protected by large national parks and reserves, geological features, archaeological and historical sites evidencing past cultures, and a great variety of ethnic groups and costumes. Still, protection of desert heritage is challenged in many countries due to low levels of security, development, and political stability. Ecotourism generates a plethora of environmental, economic, and social-cultural impacts that need to be considered when developing tourist activities in desert environments. However, ecotourism should not be envisaged as a panacea for all the challenging impacts that deserts are facing, but it can offer a sustainable tool to contribute positively to the wellbeing of people and communities living within deserts. Ultimately, desert ecotourism can contribute to the 2030 agenda of the United Nations for the Sustainable Development Goals and the conservation and management of desert heritage.

32 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Keywords sustainable development; social impacts; environmental impacts; cultural impacts; economic impacts; deserts; ecotourism definition; ecotourism potential; desert attractions; desert constraints; local communities; soft and hard ecotourists; ecotourism marketing; ecotourism definition. FCUP 33 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Introduction

Ecotourism has been suggested as a tool for the conservation and management of deserts’ natural and cultural heritage. As one of the fastest growing sectors of global tourism industry, ecotourism has a potential to serve as an environmentally, socio- culturally, and economically viable option for promoting sustainable development in the desert biome. Many factors provide opportunities to develop ecotourism in deserts, which need to be carefully considered for both nature conservation goals and local socio- economic wellbeing. For this chapter, we start by contextualizing ecotourism and its research in deserts, and then describe possible attractions and constraints exhibited by deserts to develop ecotourism. We conclude with a discussion of the potential positive and negative impacts that ecotourism may derive on different desert facets.

Brief history of ecotourism and its research on deserts

The relationship between environment and tourism has interested conservation researchers and managers for a long-time (Wagar, 1964; Budowski, 1976). The term ecotourism became popular in the 1990’s, when scholars started raising interest in the conflicts and potential solutions and symbiotic relations between tourism and the natural environment (see Ceballos-Lascurain, 1996). Multiple definitions of the term have been proposed and revised since then. In general, ecotourism refers to tourism in natural environments involving a conservation element, environmental education components, and benefit sharing with local communities (Fennell, 2001). Some scholars (Weaver, 2001a; Fennell, 2015) consider that ecotourism needs to address the following factors:

(1) be developed at small scales, for example, in typically low-density groups of travelers; (2) foster behaviors that help preserve local biodiversity and habitats, particularly by potentiating environmental awareness and learning experiences among tourists, and with local communities, the importance of biodiversity preservation through environmental education; (3) contribute economically to the preservation, maintenance, and enhancement of biodiversity and habitats; (4) respect the integrity of host communities (local beliefs and livelihoods) while committing to promote local culture and ensuring folklore maintenance (i.e., orientated to locals); 34 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

(5) generate net positive incentives to the local economies and the well-being of communities (i.e., improving the welfare of local people) sufficient to make them value and protect the surrounding wildlife heritage as an income source; (6) balance the biocentric and anthropocentric views to ensure long-term sustainability of projects; (7) be moderately commercialized.

Therefore, developing ecotourism is not a simple task, and planners, managers, and policy makers need to understand and contextualize these dimensions if they want to achieve sustainable development in tourism. Being the fastest growing segment of the tourism industry, ecotourism is estimated to generate annual revenues of more than $28.8 billion for developing countries (Kirkby et al., 2011). This is 34% more than what is spent annually by governments, private donors and official aid agencies in global conservation projects (Waldron et al., 2013), and, for example, 44% of the value needed to cover the global annual costs of protecting and managing Important Bird Areas (McCarthy et al., 2012). Yet, there are no reliable financial figures on ecotourism for specific biomes, such as deserts (UNEP, 2006a). Different types of ecotourists contribute to these expenditures and market segmentation divides them into ‘hard’ and ‘soft’ ecotourists according to their motivations, attitudes, and behaviors. Hard ecotourists are considered as having strong biocentric attitudes and commitments to environmental issues, they desire deep and meaningful interactions with the environment, travel in small groups, and search for physically challenging experiences requiring few or no tourism supporting infrastructures, while soft ecotourists exhibit the opposite traits in the ecotourism continuum. As such, the hard ecotourist may spend more time with local communities to understand their relationships with nature, whereas the soft ecotourist opts for experiences of shorter duration (Weaver, 2001a). Market segmentation is, therefore, important also for ecotourism developers as it helps to manage the natural resources and their utilization. These strategies help maximize the economic and environmental contributions of ecotourism to achieve the United Nations Sustainable Development Goals (UN-SDGs; UNWTO & UNDP, 2017), particularly in alleviating poverty, providing decent work and economic growth for all, promoting gender equality (reducing inequalities), protecting desert life, and establishing strong partnerships between the business, local communities, governments, academia, and NGOs. The desert biome is still underrepresented in the general literature of ecotourism research (Santarém and Paiva, 2015). Deserts are home to 6% of the world’s human population, mostly poor and marginalized (UNDP, 2011). They have been mostly FCUP 35 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel neglected for conservation and development funding schemes (Waldron et al., 2013). Furthermore, they contain some of the most threatened biodiversity, fragile environments, and vulnerable cultures in the world (UNEP, 2006b; Brito et al., 2014, 2016, 2018). As such, ecotourism constitutes a complementary step to preserve the natural and cultural heritage of deserts while potentially improving local livelihoods. Additional efforts are needed to design desert ecotourism products that translate into sustainable development in tourism.

Potentials and constraints of deserts to develop ecotourism

Deserts exhibit a plethora of potential attraction elements motivating tourists to visit these areas but also involve several constraints for ecotourism development (Table 1.1). Both aspects are crucial to consider, and they should take into account the pressures that ecotourism can create on the environment and the existing economic and socio-cultural uses of the desert biome (UNEP, 2006b).

Table 1.1. Potential attraction and constraining elements of the desert biome to develop ecotourism. Attraction elements Constraining elements Natural features: Security conditions Landscape Developing conditions Biodiversity Political conditions Conservation Increased accessibility Human features: Natural resource exploitation Historical Socio-cultural

Desert attraction elements for ecotourism development

Most deserts and arid regions classify as being among the last wild biomes on Earth (Watson et al., 2018), which offers excellent opportunities for ecotourists seeking ‘last- chance-to-see’ places (Lemelin et al., 2012) that are considered as wilderness areas (Saarinen, 2018). Deserts are also characterized by the high ‘visibility factor’ that increases the likelihood of successful wildlife encounters (Weaver, 2001b), including unparalleled natural sceneries and geological features, especially by ecotourists focusing on non-living natural phenomena (Weaver, 2001a).

36 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Natural features

Deserts display remarkable landscape features highly attractive for ecotourism (Fig. 1.2). They are considered natural laboratories for investigating the history of Earth, as extreme aridity and sparse vegetation provide large areas of exposed bedrock materials showing the record of geological processes operating for millions of years. Deserts provide a large number of paleontological findings (UNEP, 2006b), with the most noteworthy examples including:

• Badland terrains of the Gobi Desert, where 65 million years-old dinosaurs and mammal fossils were unraveled; • Ténéré Desert in Niger, where excavations revealed 25 tons of dinosaur fossils; • Pisco Basin in Peru, where stratigraphic sequences of Miocene to Pliocene rocks show assemblages of fossil marine mammals; • Colorado Desert in California (USA), where one of the most complete records of late Cenozoic land mammals for North America can be found.

One of the most iconic landscape features of deserts is sand dunes. Among them, the Skeleton Coast in Namibia, the 168,350 km2 of parallel dunes of the Simpson Desert in Australia, and the world highest sand dune in Atacama Desert, Chile (Cerro Medanoso at 2,100m) are the most remarkable. Sand dunes provide scenic settings that provoke strong emotional connections to nature, offer relief from high-stress daily life of urbanized people, and provide exceptional opportunities for experiencing solitude and tranquility. Other prominent landscape features of deserts are high-altitude mountains. For example, the Tibesti mountains in Chad, the Al Hajar Mountain in Oman, or the Kerman mountains in Iran all rise above 3,000 m (UNEP, 2006b). Isolated rock formations that rise abruptly from a surrounding plain, known as island-mountain or inselberg, are also highly attractive, and an iconic example is the Uluru in the Northern Territory (Australia). Indeed, deserts contain spectacular geological features of potential interest for ecotourists. A few examples from around the globe are as follows:

• Volcanic cones, such as the Trou au Natron and Emi Koussi (2,450 and 3,445 m elevation) in the Saharan Chad; • Meteorite impact craters, such as the Wabar Craters in ; • Rock canyons (deep cleft between escarpments or cliffs resulting from weathering and the erosive activity of a river over geologic timescales), such as the majestic 450-km long FCUP 37 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

and 1.5-km deep Grand Canyon (USA) or the Iherir-Imirhou valley in the Tassili n'Ajjer (Algeria); • Rock archers formed by wind and water erosion processes, such as the Aloba Arch in the Saharan Chad, the Jabal Umm Fruth Bridge in the Wadi Rum of Jordan, or the Arches National Park (USA); • Desert potholes (shallow puddles or deep pools that collect precipitation from small catchments), such as in the Colorado Plateau (USA); • Eolian yardang landforms (massive corrugated ridges), such as in the Lut Desert in Iran; • Saltpans (natural flat expanses of ground covered with salt and other minerals), such as the Chott El Djerid in the Tunisian Sahara.

Finally, wetlands are abundant in some desert areas and, given the generalized arid character of the desert landscapes, they are highly attractive to ecotourism.

38 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 1.2. Examples of landscape features typical of the desert biome with potential for ecotourism. (A and B) Sossusvlei in Namib-Naukluft National Park (Namibia); (C) Es Sba (Mauritania); (D) Guelb el Makhsar (Mauritania); (E) Tassili n'Ajjer (Algeria); (F, G and H) Tadrart (Algeria). Courtesy of José Carlos Brito and Jarkko Saarinen.

FCUP 39 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Deserts contain biodiversity features that can be highly attractive for ecotourism. These include endemic species exhibiting unique adaptations to aridity that are found in no other biomes (Fig. 1.3; Weaver, 2001b). Examples of extreme adaptations and uniqueness's found in deserts include:

• Namib Welwitschia (Welwitschia mirabilis), among the most ancient organisms in the Plant kingdom and considered a living fossil; • Southwestern USA Joshua tree (Yucca brevifolia), the symbol of the local National Park that displays a very specialized mutualistic pollination system with the yucca moth (Megathymus sp.), which spreads pollen while laying eggs inside the flower in a communal way; • Fringe-fingered ( spp.) that perform sand-swimming to escape predators in the Sahara Desert; • Australian thorny devil (Moloch horridus) that displays a unique water transportation system through its ridged thorny scales that enable the animal to collect water from simply touching it.

These remarkable adaptations to arid conditions have been used to market them as flagship species, that is, species that due to their morphological, behavioral, and cultural traits can be used as the focus of broader conservation campaigns and can be distinctly promoted according to specific market segment traits. For example, soft ecotourists may prefer large and easily visible species, while hard ecotourists may opt for ‘last- chance- to-see’ endemic endangered species (Lemelin et al. 2012; Santarém et al. 2019). Deserts contain also some of the most endangered species in the World, which may prompt ecotourists to pay high fees to see them in the wild and, thus, contribute financially to their preservation. In addition, wetlands in deserts constitute local biodiversity hotspots supporting particular fauna suitable for ecotourism purposes. For example, the rock-pools of Mauritanian Sahara gather relict populations of West African crocodiles (Crocodylus suchus; Brito et al., 2014) and the wetlands near the archaeological city of Palmyra (Syria) support a relict colony of the critically endangered Northern Bald Ibis (Geronticus eremita; Serra, 2007), thus offering a great opportunity for desert wildlife viewing.

40 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 1.3. Examples of biodiversity features in the desert biome with potential for ecotourism: (A) African golden wolf (Canis anthus); (B) Felou gundi (Felovia vae) and Rock agama (Agama boulengeri); (C) Desert elephant (Loxodonta africana); (D) Namibian giraffe (Giraffa camelopardalis angolensis); (E) Greater hoopoe-lark (Alaemon alaudipes); (F) Brown- necked Raven (Corvus ruficollis); (G) Lanner falcon (Falco biarmicus); and (H) Desert monitor (Varanus griseus). Courtesy of Frederico Santarém and Jarkko Saarinen. FCUP 41 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Deserts contain conservation features that improve the chances of wildlife encounters, which enrich the recreational experience for ecotourists. Unique biodiversity is generally safeguarded by a network of protected areas. For instance, the Aïr and Ténéré Addax Sanctuary and the Termit-Tintoumma Natural and Cultural Reserve in Niger were designated to protect the critically endangered and flagship species Addax (Addax nasomaculatus), and the Joshua Tree National Park (USA) was established for protecting the native Joshua trees (UNEP, 2006b). Protection of biodiversity and ecological processes is strongest in North American and Australian deserts, where the Grand Canyon (1919) and Simpson Desert (1967) National Parks were established a relatively long time ago. On the contrary, the desert areas of developing countries are still under protected (Brito et al., 2016; Weaver, 2001b). Multiple protected natural sites displaying outstanding evolutionary and landscape features have been recognized even in the most remote parts of deserts, which appeal to a specialized segment of ecotourists that are able to travel long distances to observe unique desert features (hard-desert ecotourists; Santarém et al., 2018).

Human features

Numerous features associated with the historical human occupation of deserts make them attractive to ecotourism (Fig. 1.4). These include rock art found in caves, mostly depicted as paintings and engravings, illustrating past livelihoods and even past climatic shifts. The World Heritage site of Tassili N’Ajjer in Algeria and the Messak Settafet in Libya are considered open books of History, with thousands of rock paintings and engravings catalogued. Rock art is highly recognizable among international ecotourists because in deserts it is easy to find it in large concentrations, offering abundant possibilities to interpret how humans interacted with the environment and how they developed associated spiritual and community values, making rock art in the desert biome among the best sources of human history. The low disturbance of desert soils preserved remarkable archeological remains and help to understand how past societies lived in arid environments. Ruins, tombs and other sites of historical land occupation highlight how successive human societies shaped the desert environment (UNEP, 2006a; Serra, 2007). For example:

• The earliest run-off irrigation system supporting large cultivated fields in the world was discovered in the Negev Desert (Israel); 42 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

• The ancient city of Palmyra in Syria – where for centuries nomadic Bedouins traded livestock and relative products – has been revealing artifacts of world recognized value; • Tombs in Egypt show how people worshiped iconic desert species such as the crocodile, represented in art linked to Sobek, god of Nile.

The historical crossing of deserts along trading routes allowed the establishment of trading villages, acting as commercial warehouses and resting places. For instance, the old salt caravans in the Sahara or the Silk Road in Central Asia crossed desert areas through established routes with stopovers in caravan villages. Some of these villages are now classified as World Heritage Sites and offer extensive opportunities to explore desert history (UNESCO, 2018), such as Tombouctou (Mali) and Chinguetti (Mauritania) where extensive trading network enabled the establishment of ancient libraries of notable historical importance. FCUP 43 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 1.4. Examples of historical features associated with human occupation of the desert biome with potential for ecotourism: (A) rock-art representation of giraffe (Tadrart; Algeria); (B) rock-art representation of crocodile (Messak Settafet; Libya); (C) World Heritage Site of Ghadames oasis (Libya), (D) Oualata and (E) Tichit (Mauritania), ancient stop- points for caravans crossing the Sahara; (F) historical library in Chinguetti (Mauritania).

Deserts display several cultural features associated with human presence that add value to the general recreational experience of the ecotourist and enhance strong connections between dissimilar societies (Fig. 1.5). Firstly, desert people domesticated camels and dromedaries due to their adapted physiology to withstand hyper-arid conditions and to transport goods along extensive trade routes. Also, humans have long adapted to the harsh environment by using desert elements to counter the high temperatures, strong winds and lack of water (Saarinen, 2016). For instance: 44 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

• Tuareg tribes in the Central Sahara use dried bulrush (Typha elephantina) to build the roofs of their houses (locally called Zeribas) and transport water and milk in reservoirs made of goat skin (called Guerba) that keep the water fresh; • Moors in the West Sahara use dromedary fur and local trees to build their tents (called khaimas); • OvaHimba people (considered the last nomadic people of Namibia) cover their bodies with a self-prepared cosmetic mixture containing an ochre pigment that protects them from the extreme dryness and gives their skin a red characteristic; • Wool hand-made crafts are manufactured in Egypt for millennia and are now highly appreciated by foreigners looking for desert handicrafts.

Secondly, due to the unpredictability of rain in deserts, humans shaped the environment to benefit their livelihoods and prosper in less favorable climatic times. The best examples of these modifications in the arid environment are the oasis, fertile green areas created for agriculture (mostly date palms, Phoenix dactylifera) and whose location has been of critical importance for trading and transportation routes because of the provided shade in the hyper-arid landscape. Oases become fertile when freshwater is extracted on the land surface or underground using man-made structures. Examples include:

• The extensive rock-walled terraces built by the Trincheras people of northern Sonora desert (USA) (surface extraction); • The qanats, networks of subterranean channels developed by the Persian in Iran and now used throughout the North African and Asian deserts (underground extraction).

In the region of Gourara and Touat in the Algerian Sahara, about 500 of these systems are still working. Where water-bodies are far from the oases, as in much of the Sahara, the water is lifted to the surface by cantilevered bucket systems (called shadufs). Oases vary in extension, but one of the largest and most spectacular is in Uzbekistan, which supports 230,000 ha of irrigated land. Thirdly, early hunter-gatherers developed uses for local plants and animals to enrich their medicine and diets. For instance, the Topnaar people of the Namib Desert have known uses for 81 plant species and the San people of the Kalahari Desert list more than 100 species of food plants and 55 animal species, some of them with medical uses. There are unique ethnic costumes and ceremonies that can only be appreciated in deserts. For example: FCUP 45 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

• Religious ceremonies are pursued by the Shoshone people around the Yucca Mountain (USA) and by the aboriginal people around Uluru (Australia); • Assyrians, indigenous to Western Asia, are well known by the khigga, a folk dance that is traditionally performed during wedding ceremonies.

All modern people living in deserts represent a great repository of knowledge concerning the life in harsh environments, which must be protected and acknowledged by specialized ecotourists seeking to deep their knowledge on local desert cultures (UNEP, 2006a,b; Weaver, 2001a). 46 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 1.5. Examples of cultural features associated with human occupation of the desert biome with potential for ecotourism: (A) settlement and food-storage systems of OvaHimba people in Opuwo (Namibia); (B) settlement in Ounianga Serir (Chad); (C) oasis (Ziz valley, Morocco); D - riverbank village (Tassili n'Ajjer, Algeria); (E) desert village (Agadez, Niger); (F) trading route (Aïr mountains, Niger); (G) Watering livestock from rock-pool; and (H) excavated well from dry riverbed (Tagant, Mauritania). FCUP 47 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Constraints to ecotourism development

Desert environments involve certain features that may constraint the development potential of ecotourism. Some may affect particular deserts, such as armed conflicts or poor development conditions limiting accessibility. Others limiting issues may refer to common trends across world deserts, such as increased accessibility and exploitation of natural resources. In all cases, these constraints diminish some of the most valuable desert assets to ecotourists (Santarém & Paiva, 2015; Brito et al. 2016). However, hard and soft ecotourists may perceive these constraints differently and may be prompting to travel or not, respectively, to desert regions where these constraints critically shape desert ecotourism attractiveness (Santarém et al., 2018). For instance, hard ecotourists may support the idea of travelling to remote desert regions of poor accessibility, lacking the comforts provided by regular infrastructures, while soft ecotourism may benefit from the same transportation system created by mining or other extractive industry development purposes (Weaver, 2001a).

Security conditions

Poor security conditions diminish the attractiveness of some deserts to ecotourists, who may feel unsecure to travel to regions facing on-going, and, in some cases, long-term conflicts. For example, warfare in North Africa, the Middle East, and in Southwestern Asia has caused environmental destruction (e.g., first Gulf War) and the massacre of local people, migrants and travelers (e.g. Darfur conflict). Conflicts in the Sahara-Sahel are contributing to the extirpation of threatened populations of flagship desert species at the fastest rate ever (Brito et al., 2018). Additionally, conflicts threaten cultural sites of global recognizable importance, such as the historical libraries in Tombouctou (Mali) or the archaeological city of Palmyra (Syria). The depletion of natural and cultural values threatens the long-term sustainability of local ecosystems to support any form of tourism (Serra, 2007). Furthermore, the attacks against civilians and, specifically, the kidnapping for hostage taking counteracts against the countries where attacks are perpetrated. For instance, the Ministries of Foreign Affairs of developed countries generally issue advises against travelling to such countries (or define ‘no-go areas’ within countries) as response to protect their national travelers. Another negative aspect of hostage taking for future ecotourism development is the long-lasting reputation of danger associated to the region and the psychological burden imposed by traveling in those regions (even after conditions ameliorate). 48 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Under-developed conditions

Large portions of the world deserts span countries with poorly developed conditions (UNEP, 2006b). In general, desert regions in North and South America, in Australia, and in Southern African countries are perceived as socio-economically developed and have relatively good infrastructure, whereas deserts in North Africa, the Middle East, and Southwestern Asia display poorer social conditions and infrastructures. Underdeveloped airports, hazardous roads, lack of regular transportation networks, alongside the lack of supporting infrastructures (e.g., hospitals, accommodation, restaurants) may affect the attractiveness of deserts for ecotourists. For example, central Saharan countries, such as Niger and Chad, are considered among the least infrastructure-developed countries in the world, whereas Namibia is recognized to be among the top countries in developed tourism infrastructures, specifically in the desert biome. Although deserts experience less disease outbreaks in comparison to other world biomes, ecotourists may get severely ill from local epidemics, such as the Arizona (USA) endemics Hantavirus Infection, the Rocky Mountain Spotted Fever, and the Coccidioidomycosis Valley Fever (AZDO, 2019).

Political conditions

Ecotourists may express prejudice against visiting deserts located in countries with poor levels of democratic functioning or under sanctions by the international community. The political agendas of several countries in the Sahara-Sahel, Arabian Peninsula and Southwestern Asian deserts are not always aligned with the interests of both local and international communities. In countries ruled by dictatorial regimes, where freedom of thought and speech are often suppressed, animosity between locals and ecotourists may generate, which negatively affects the promotion of desert ecotourism and the attractiveness of particular regions (Brito et al., 2014, 2016). Furthermore, the high levels of bureaucracy and corruption that usually characterizes dictatorial regimes hamper the establishment of successful ecotourism businesses.

Increasing accessibility

Accessibility plays a critical role in attracting soft and hard ecotourists to deserts but the assessment of its suitability to attract both groups remains poorly studied. On the one hand, increasing accessibility promotes the visit of soft ecotourists, but on the other FCUP 49 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel hand it will tend to deter the hard ecotourists (see Figure 8 in Chapter 2: Desert biodiversity – world’s hot spots/globally outstanding biodiverse deserts; Santarém, et al., 2018). In addition, human population is growing worldwide, and desert areas are no exception; urbanization is increasing, and nomads are becoming sedentary in many world deserts (UNEP, 2006b). These developments cause generalized landscape degradation, with cascading negative effects upon large-sized animals needing vast wild areas during seasonal migratory movements. For example, migratory populations of the flagship African savannah elephant (Loxodonta africana) in Mali have been particularly affected by increasing human sedimentary in elephant’s migratory pathways (Brito et al., 2018). Land degradation and biodiversity loss lessen ecotourists expenditures as desert natural features disappear.

Exploitation of natural resources

The exploitation of natural resources in world deserts, including oil and gas pumping and minerals extraction is expanding. Besides deteriorating habitats in which ecotourism operations are based, they also affect the ecotourists’ experience (Weaver, 2001a). For example, oil exploration activities have increasingly affected populations of the critically endangered Addax in Niger. This flagship species has been experiencing dramatic declines since oil exploitation activities started (Brito et al., 2018), which prevent ecotourists from travelling there, as the most important ecotourism asset may be completely depleted from the ecosystem.

Impacts of ecotourism on different desert facets

Resolutions adopted by the General Assembly of the United Nations highlighted the positive impacts of ecotourism to eradicate poverty and to protect the environment (UNWTO, 2018). Ecotourism has a potential to contribute to these aims in desert environments (Table 1.2) but it is important to realize that this ‘potential’ can be positive or negative and direct or indirect. Positive impacts tend to occur at broad scales, whereas negative impacts are more localized. Developing desert ecotourism should also consider these aspects to maximize return-on-investment scenarios at multiple scales (from the business owner to local communities passing through state governments) and to diminish economic inefficiencies in peripheral regions, while preserving local environments at the same time. Despite these potential negative impacts, ecotourism 50 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

has been suggested to be among the best tools to address economic and social imbalances in deserts (Santarém and Paiva, 2015).

Table 1.2. Potentially positive and negative impacts from ecotourism in desert biome. Positive Negative Environmental impacts • Protects species and habitats • Disturbs or destroys wildlife • Repairs degraded habitats • Potentiates species invasiveness • Funds protected areas • Restructures habitats permanently • Maintains ecosystem services • Generates local pollution and waste • Subsidizes scientific research • Boosts water abstraction • Raises conservation awareness and • Diverts funds from less to more fosters environmental-friendly behaviors charismatic species

Economic impacts • Generates direct/indirect employment • Needs substantial initial investments and revenues and operational costs linked to remoteness and climate • Multiplies systems of generating • Sensitivity to long-term uncertainties in revenues generating revenues • Funds community-based projects • Increases medical assistance costs • Stimulates other forms of tourism • Disrupts local subsistence • Supports pharmaceutical investigation • Creates income inequalities and increases economic pressures

Socio-cultural impacts • Improves education • Increases cultural shocks • Diminishes gender inequalities • Erodes local power • Sustains ancient practices • Generates animosity/resentment • Provides spiritual benefits • Fosters social inclusiveness

Modified from Weaver, D. B. (2001) Ecotourism. Queensland, Australia: John Wiley & Sons Australia, Ltd.

FCUP 51 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Ecological impacts

Positive Ecotourism may protect desert biodiversity through financial mechanisms that support breeding, feeding, translocation, anti-poaching, and other conservation programs with direct positive effects on species population growth and viability (Buckley et al. 2016). For instance, the populations of cheetah (Acinonyx jubatus) in the Kgalagadi Transfrontier Park (South Africa) and of the Western Barred Bandicoot (Perameles bougainville) in the Arid Recovery Reserve (Australia) rely on tourism revenues for their conservation. Ecotourism may also benefit biodiversity and human populations when it contributes to the restoration of degraded habitats and environments that help to maintain and safeguard ecosystem services (Buckley, Morrison, & Castley, 2016; Weaver, 2001a). These include provision of clean water, which in deserts is a scarce and highly valuable asset, and raw materials, such as palm trees for timber production, and medical resources, such as venoms from desert vipers for anti-venom development (Weaver, 2001a). Ecotourism may fund the establishment of new protected areas in desert countries and maintain and increase old ones. For instance, 12% of the Namibian protected areas’ budgets are assured by tourism expenditures (Buckley et al. 2012). Ecotourism may also subsidize scientific research on species and protected areas (Wearing, 2001), by making people donating more funds to environmental causes and creating “environmental watchdogs” (nature activists that will intervene in favor of the environment; Weaver, 2001a). For example, the conservation of the northernmost population of the African savannah elephant in the Gourma region (Mali) is enforced by a community-based vigilance network composed by 520 “eco-guardians” living throughout the elephant range, which continually patrol and alert about elephant locations, numbers and behavior, elephant deaths, and poaching incidents (Brito et al., 2018). By raising conservation awareness and fostering environmental-friendly attitudes among local communities, policy makers and the likes, ecotourism may influence communities’ nature-pro behaviors, improving the sustainable harvesting of food and natural resources (Lindberg, 2001; Wearing, 2001; Weaver, 2001a).

Negative Recreational activities in deserts are usually based in all-terrain vehicles, which may degrade the landscape and increase local pollution. The exhaust gases released to the atmosphere may also affect sensitive desert vegetation. For example, many plants of the Mojave Desert (USA) died when exposed to various vehicles’ pollutants in a laboratory experiment. The impacts of chemical fuel leakage to the ground and of soil 52 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

compacting and erosion may take decades to disappear. For instance, the soils of the Joshua Tree National Park (USA) were found to be compacted from 2 to 5 cm because of constant vehicles traffic (Ouren et al., 2007). Motorized driving may also run-over desert fauna or destroy vertebrate ground nests. For example, the populations of desert tortoise (Gopherus agassizii) and Couch’s spadefoot toad (Scaphiopus couchii) have decreased in areas with high off-road vehicles passage in Mojave and Colorado deserts (USA) because vehicles collapsed burrows. In addition, they may produce noise that disturbs local fauna, which ultimately alters animal behavior, reduces preys’ capabilities of detecting predators, and stimulates aestivating animals to emerge from their burrows at inappropriate times. For instance, kangaroo rats (Dipodomys deserti) in California deserts (USA) experienced hearing loss when exposed to continuous sound produced by dune buggies, which made them unresponsive to predator sounds for weeks (Ouren et al. 2007). Additionally, desert vegetation is under increased risk of vehicles passage, being reported that lichen communities of the coastal Namib desert are eliminated by a single vehicle passage (UNEP, 2006b). Transportation can contribute to the proliferation of invasive species in deserts, particularly when humans assist plant seeds or invertebrates to travel long distances, for instance attached to vehicles. Proliferation of invasive species may also be facilitated by vehicles’ exhausting gases. For example, annual grasses prone to wildfires proliferated

after a CO2 increase in the atmosphere of the Mojave Desert (USA), accelerating local fire frequency in regions where they did not exist before (UNEP, 2006b). Keeping vehicle speeds down and, more importantly, using other forms of transportation, such as camel trekking or air ballooning, can reduce air and soil pollution, preserve the quality of the surrounding environment, and minimize wildlife direct damages (UNEP, 2006a). Permanent environmental restructuring may occur when setting and operating tourism-supporting infrastructures. Locally, the surrounding landscapes may become less natural, sound and light pollution may become more frequent, and the soils sustaining those infrastructures may become permanently altered. In the long-term, ecotourism may also generate waste and contaminate soils and the scarce water resources available, if wastes and sewage waters are not properly treated (Serra, 2007; UNEP, 2006a). As ecotourism potentiates the development of other economic activities, additional infrastructures may be constructed, which can magnify impacts on soil and water, particularly when less-benign energy sources are used (Weaver, 2001a). Several measures should be taken to minimize the ecological impacts of ecotourism. For example, construction of infrastructures should opt for local materials to minimize landscape impacts. Alternative energies should be preferred when building/recovering infrastructures, such as wind and solar energy that provide feasible FCUP 53 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel energy production in deserts. Ecotourists should be encouraged to take toxic wastes (like batteries and aerosols) back home and arrangements should be enforced to deal with waste removal. For example, companies from Mauritania and France set-up a joint waste collection protocol in Chinguetti, a highly visited UNESCO World Heritage Site (UNEP, 2006a). Water abstraction for serving ecotourism activities and accommodation may cause water shortages. This problem is amplified by the natural dryness of deserts. For example, the small dune fields throughout the Taffilat region of Morocco are threatened by the massive and unsustainable construction of tourism facilities, which are progressively surrounding them. Making use of systems that reduce water-levels consumption and recycle sewage waters should be preferred whenever possible. Raising awareness among ecotourists about the consequences of their actions should stress the scarcity of water resources and the unsustainable nature of its excessive use in desert environments (UNEP, 2006a). To a lesser extent, the promotion of the most charismatic desert species by ecotourism operators may counteract against other species less appealing to the general public, diverting funds from the latter to the former. This may have a negative impact on less charismatic desert species when international ecotourists restrain from conserving them (Weaver, 2001a). In alternative, the promotion of flagship fleets, i.e. groups of species displaying similar appealing traits that raise the profile of more than one species to attract different segments of the society, may maximize conservation marketing and return-on-investment scenarios for ecotourism in deserts (Santarém, et al. 2019).

Economic impacts

Positive Ecotourism can provide direct and indirect economic benefits to desert communities, offering local employment as tour guides, receptionists, cookers, planners etc. Some jobs are desert-specific and benefit from local people’s in-depth knowledge of the environment, such as camel drivers, who become indispensable to convey desert cultural knowledge to ecotourists. Ecotourism may also provide alternative ways of empowering local communities, for instance by creating opportunities for selling desert crafts to travelers (UNEP, 2006a), further beneficiating the whole community. Ecotourism may stimulate other forms of tourism, such as cultural tourism or adventure tourism, and act also as a catalyzing agent to generate other jobs that feed it directly or indirectly. By creating employment, ecotourism displays an underlying multiplier effect that benefit local communities and regional and national governments (Wearing, 2001; Weaver, 54 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

2001a). For instance, it may help creating jobs in provisioning (such as food production or bed linen) and services sectors (such as water cleaning, infrastructure maintenance or banking and accounting). These indirect effects occur, for instance, when local restaurants use ecotourists’ expenditures to purchase local goods and services from other businesses, and to pay employees that will spend their wages in buying external goods and services. This generates a circular economy where all parts benefit. Additionally, all these sectors must pay taxes, fees, and licenses that generate increased stimulus to fiscal policies (Lindberg, 2001). Peripheral and marginalized regions of some world deserts should be the ones to benefit the most, which labels ecotourism as a sustainable tool to eradicate extreme poverty (Lindberg, 2001; UNWTO, 2018; Weaver, 2001a). For instance, local people generated multiple jobs (directly and indirectly related to the tourism sector) after successive Communal Conservancies were created in Namibia. Residents of ecotourism destinations may benefit from revenue-sharing programs that fund community-based projects, such as sustainable agriculture or clinic and school construction and upgrade. For example, ecotourism helped the Torra Conservancy of Wilderness Safaris’ Damaraland Camp in Namibia to release c.$120,000 of the profits to upgrade local schools in 2002 (Butler et al. 2009). Ecotourism may also increase the investigation and procurement of desert natural assets for food production, such as palm dates, or pharmaceutical developments, such as desert venoms, which benefit both local and global communities (Weaver, 2001a).

Negative Initial investments on land acquisition, materials’ purchases (including vehicles), initial staff training, and promotion have substantial costs, but on-going operational costs, such as wages, infrastructures and vehicles maintenance (meaningful in deserts due to the remoteness and harsh climate of the biome – temperature, sand and salt winds), and continuous marketing bail out most of the operational costs when developing desert ecotourism (Serra, 2007). Ironically, these costs translate into benefits when considered from the perspective of the people that benefit from them (Weaver, 2001a). Building infrastructures nearer established cities, villages and oases, and opting for vehicles with anticorrosive materials may diminish the depreciation speed of these goods. Ecotourism development in the desert biome may be sensitive to long-term uncertainties in generating revenues. These may arrive from the general supply-and- demand risks but also from specific traits of some deserts, such as regional insecurity, or diseases outbreaks. These aspects may make initial investments unreliable, resulting in no-profit situations or, even worse, in substantial economic losses (Brito et al., 2014, FCUP 55 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

2016; Weaver, 2001a). Ceasing regional conflicts and developing control measures to prevent disease outbreaks helps to re-establish operations. For instance, after a 10-year break period following the growing in regional insecurity, only recently international tourist agencies are again offering services to visit some Sahara-Sahel countries. That is the case of Mauritania, where French carriers re-established flights to well-known touristic places in the Adrar and areas. Ecotourism may incur in indirect economic costs when wildlife encounter accidents occur, particularly when flagship species protected for ecotourism become cause of human injuries. For example, the venomous desert American horned rattlesnake (Crotalus cerastes) and the North African horned viper (Cerastes cerastes) are among the world’s most toxic and they may bite visiting ecotourists, generating additional costs related to medical assistance, notwithstanding the animosity that injured people create towards typical desert flagship species (Weaver, 2001a). Developing clinics and hospitals in strategic places would allow injurers to be healed faster and properly, reducing the costs of transportation of diseased people. Multinational pharmaceuticals can also target efforts to develop strong antidotes specific to each of the several desert venom species, which entail substantial costs in research and development but translate into profits for the world society at many scales in the long term. Especially in the Global South, ecotourism operations are often owned and monopolized by international business and foreign actors controlling all processes from the land acquisition to the establishment of an ecotourism company, including the provision of related services. Considerable negative impacts might arise if foreign control and importation of external products leak all the revenues away from the local community and the regional states where ecotourism takes place (Lindberg 2001; Weaver, 2001a). Including local communities in early strategic ecotourism development plans and developing strong policies against monopolies and expropriation of local deserts from foreigners enforces local people rights and grants that most of the revenues stay with who need them the most. Corporate social responsibility is a recent type of business self- regulation that can help on this issue, as it includes men, women and the youth in strategic plans to increase long-term companies’ profits. Ecotourism, as any other form of tourism, may sometimes exacerbate income inequalities within community members (Lindberg, 2001). Ecotourism may counteract against local interests and preferences by increasing demand-to-supply and the prices of house rents, local goods and services, which create economic pressures that may disrupt people financial stability and obliges them to leave the region, and may extirpate local ancient cultures (Lindberg, 2001; Wearing, 2001; Weaver, 2001a). In Mmatshumu 56 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

village of the Makgadikgadi pans landscape (Botswana), for example, some communities have become involved in ecotourism, which has resulted in a decline or even termination in traditional livelihood activities like subsistence hunting and gathering. However, not all community members have been able or willing to work with ecotourism and related employment, which has caused increasing inequalities in communities (Lenao etl al., 2014). To avoid ecotourism related local inequalities, strong national policies promoting equitable access to resources and considering the needs of local people may boost the development of desert communities. Additionally, alternative jobs should be considered in remote regions where few desert attractions for ecotourism exist, thus allowing people to sustain the same lifestyle as those living in ecotourism-rich regions. Regional agreements and international conventions must develop strong legal frameworks for the protection of desert natural assets and the interests of indigenous people.

Sociocultural impacts

Positive Successful ecotourism development can diminish gender inequalities and give power to women and youth embracing qualified jobs. It has the potential to improve education in peripheral and marginalized desert regions and to generate resources for community-based development projects, improving people’s acceptance of alternative ways to generate income while preserving desert resources (UNWTO, 2018). For instance, nomadic Bedouins that once hunted birds in Palmyra (Syria) adopted more sustainable activities by becoming eco-guides (Serra, 2007). Ecotourism creates markets for local products, thereby sustaining and promoting ancient practices, and making local people proud of their traditions. By turning local knowledge, cultural skills and traditional craft skills into attractions to travelers, ecotourism may help to preserve the desert cultural heritage. For example, the Ksour Route is a project developed to show ecotourists the old trade routes across the Southern Algerian Sahara and that aims to rehabilitate traditional architectural heritage. A spectacular example comes from the La Routa de Sonora Ecotourism Association, which is a transboundary commitment between USA and Mexico to protect the natural and cultural heritage of the Sonora desert through desert trans-frontier ecotourism (UNEP, 2006a). Ecotourists that spend more time understanding host communities’ habits and patterns will likely respect local knowledge and traditions and promote them in their own countries, creating a society ruled by respectful understanding of social differences (Wearing, 2001).

FCUP 57 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 1.6. An example of successful cooperation between local communities, research institutions, national park staff and governmental institutions in the promotion of environmental responsible acts in deserts and in early adoption of ecotourism as a means to preserve the West-African crocodile (Crocodylus suchus) in Diawling National Park, Mauritania. Courtesy of Joana Marques. Erro! Não existe nenhum texto com o estilo especificado no documento.

Ecotourism may provide many spiritual benefits to the individual, maximizing nature enjoyment for locals and ecotourists alike (Weaver, 2001a). Deserts are particularly prone to this spiritual relief sensation, given their extensive landscapes that offer a perception of solitude and deepens root-connections with nature. Ecotourism may also potentiate social inclusiveness by accommodating a wide spectrum of the society, particularly individuals with substantial physical and mental limitations that could find barriers to their enjoyment in other types of activities (UNWTO, 2018; Weaver, 2001a). Deserts are especially suitable to accommodate disabled people, as their extensive plains and sparse vegetation may facilitate locomotion.

Negative While having a great potential to benefit communities, ecotourism operations can generate cultural conflicts and shocks between ecotourists and indigenous communities when certain aspects of local lifestyle are in opposition with principles defended by ecotourists (e.g. brutal killing of a charismatic animal for ecotourists that is at the same time a valuable food for local people; Weaver, 2001a). To prevent conflicts and cultural 58 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

shocks, thematic travel arrangements should respect desert traditions and be made only after consulting desert local communities (UNEP, 2006a). Marketing ecotourism should make aware the local traditions and potential conflicts of interest to visitors when visiting particular desert regions. Ecotourism may transform and erode local power systems and structures when new business-oriented local or non-local elites take control over local assets. Businesses that impose control over local desert assets, like food and water, may create resentment among local people. Also, disruption of established social relationships may potentiate local resentment upon nature that is protected for ecotourism, particularly when local access to natural assets is blocked due to stringent nature protection laws but allowed to ecotourists visiting the region. Deliberate acts against nature (e.g., intentional wild killing and habitat destruction) may occur to intentionally shock the visitors (Lindberg, 2001; Serra, 2007; Wearing, 2001). Fortunately, there are few or no reported records of these behaviors in deserts to date. Ecotourism may be seen as a complementary tool to other local practiced land-uses, and not a mean to the obstruction of ancient practices. Therefore, access to local resources should not be exclusively granted to foreigners or passively prohibited to locals. Ecotourism may generate resentment among people from the same community, when some or all of them are not involved in the ecotourism planning process, or when locals feel their visions and traditions are being neglected. Even worse, foreign-owners of ecotourism businesses may impose alien values (e.g. globalism and Western modernization) that can erode local ones and may lead to socio-cultural tensions. The involvement of local communities is one of the most critical components in determining the success of local support for nature conservation and desert ecotourism development (Figure 1.6; Serra, 2007; Wearing, 2001). Inclusive local participation in ecotourism development through corporate social responsibility initiatives can help to establish scenarios that benefit all ecotourism stakeholders without jeopardizing local values (UNWTO, 2018). To achieve the best standards of social fairness and equity it is required to consider the inherent specificities of each desert community when developing ecotourism. An example of good practice in action came from Uluru Kata Tjuta National Park (Australia), where desert Aboriginal people operated a tour company and managed all the activities, including the interpretation of Aboriginal rock-art, medical plants, and bush cookery (UNEP, 2006a).

FCUP 59 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Conclusions

The desert biome contains a wide spectrum of potentials and constrains for ecotourism planning, development, and management. On the other hand, ecotourism may deliver different levels and extents of environmental, economic, and socio-cultural impacts across global deserts. To have successful ecotourism operations, the diversity of potential impacts needs to be understood and proactively considered. Particularly because they may contribute to the increase or the depletion of desert natural and cultural features appreciated by ecotourists. We suggest that the various impacts of ecotourism are managed in respects with the specificities of each desert context, considering its people, wildlife, and ecosystems. In the desert biome, ecotourism should be developed in ways that are based on effective management tools and responsible practices leading to sustainable governance and use of the natural and cultural heritage resources of deserts. Additional research efforts are highly needed in this regard, particularly in what refers the studying of alternative and sustainable development methods. Although ecotourism is not a panacea for widespread biodiversity losses, if properly managed, it can relief some of the threats to desert biomes. Ultimately, desert ecotourism can act as a sustainable tool to achieve the Sustainable Development Goals of the United Nations, contributing positively to the wellbeing of people living within deserts and to threatened desert heritage.

60 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

References Arizona Board of Osteopathic Examiners in Medicine and Surgery (AZDO) (2019) An Introduction to Arizona’s Infectious Diseases. Accessed on 14 January at https://www.azdo.gov/. Brito, J. C., Durant, S. N., Pettorelli, N., Newby, J., Canney, S., Algadafi, W., Rabeil, T., Crochet, P.-A., Pleguezuelos, J. M., Wacher, T., de Smet, K., Gonçalves, D. V., Ferreira da Silva, M. J., Martínez-Freiría, F., Abáigar, T., Campos, J. C., Comizzoli, P., Fahd, S., Fellous, A., Malam Garba, H. H., Hamidou, D., Harouna, A., Hatcha, M. S., Nagy, A., Silva, T. L., Sow, A. S., Vale, C. G., Boratyński, Z., Rebelo, H. & Carvalho, S. B. (2018). Armed conflicts and wildlife decline: challenges and recommendations for effective conservation policy in the Sahara-Sahel. Conservation Letters 11, e12446. Brito, J. C., Godinho, R., Martínez-Freiría, F., Pleguezuelos, J. M., Rebelo, H., Santos, X., Vale, C. G., Velo-Antón, G., Boratyński, Z., Carvalho, S. B., Ferreira, S., Gonçalves, D. V., Silva, T. L., Tarroso, P., Campos, J. C., Leite, J. V., Nogueira, J., Álvares, F., Sillero, N., Sow, A. S., Fahd, S., Crochet, P.-A. & Carranza, S. (2014). Unravelling biodiversity, evolution and threats to conservation in the Sahara-Sahel. Biological Reviews 89, 215-231. Brito, J. C., Tarroso, P., Vale, C. G., Martínez-Freiría, F., Boratyński, Z., Campos, J. C., Ferreira, S., Godinho, R., Gonçalves, D. V., Leite, J. V., Lima, V. O., Pereira, P., Santos, X., Ferreira da Silva, M. J., Silva, T. L., Velo-Antón, G., Veríssimo, J., Crochet, P.-A., Pleguezuelos, J. M. & Carvalho, S.B. (2016). Conservation biogeography of the Sahara-Sahel: additional protected areas are needed to secure unique biodiversity. Diversity and Distributions 22, 371-384. Buckley, R., Castley, J. G., Pegas, F. de V., Mossaz, A. C., & Steven, R. (2012). A population accounting approach to assess tourism contributions to conservation of IUCN-redlisted mammal species. PLoS ONE 7, e44134. Buckley, R., Morrison, C., & Castley, J. G. (2016). Net effects of ecotourism on threatened species survival. PLoS ONE 11, e0147988. Budowski, G. (1976). Tourism and Environmental Conservation: Conflict, Coexistence, or Symbiosis? Environmental Conservation 3, 27-31. Butler, K., Kinneman, G., & Morin, S. (2009). The contribution of ecotourism in local development: The case of the Torra Conservancy. Interactive Qualifying Project Report. Worcester, MA: Worcester Polytechnic Institute. Ceballos-Lascurain, H. (1996). Tourism, ecotourism and protected areas. IUCN (World Conservation Union). Switzerland: Gland. FCUP 61 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fennell, D. A. (2001). A content analysis of ecotourism definitions. Current Issues in Tourism 4, 403-421. Fennell, D. A. (2015). Ecotourism (4th ed.). London, UK: Routledge. Kirkby, C. A., Giudice, R., Day, B., Turner, K., Soares-Filho, B. S., Oliveira-Rodrigues, H., & Yu, D. W. (2011). Closing the ecotourism-conservation loop in the Peruvian Amazon. Environmental Conservation 38, 6-17. Lemelin, H., Dawson, J., & Stewart, E. J. (2012). Last-chance tourism: Adapting tourism opportunities in a changing world. London, UK: Routledge. Lenao, M., Mbaiwa, J., & Saarinen, J. (2014). Community Expectations from Rural Tourism Development at Lekhubu Island, Botswana. Tourism Review International 17, 223–236. Lindberg, K. (2001). Economic Impacts. In Weaver, D. (Ed.). The Encyclopedia of Ecotourism. Wallingford, UK: CAB International. McCarthy, D. P., Donald, P. F., Scharlemann, J. P., Buchachan G. M., Balmford, A., Green J. M., Bennun, L. A., Burgess, N. D., Fishpool, L. D., Garnett, S. T., Leonard, D. L., Maloney, R. F., Morling, P., Schaefer, H. M., Symes, A., Wiedenfeld, D. A., & Butchart, S. H. (2012). Financial costs of meeting global biodiversity conservation targets: Current spending and unmet needs. Science 338, 946-949. Ouren, D. S., Haas, C., Melcher, C. P., Stewart, S. C., Ponds, P. D., Sexton, N., R., Burris, L., Fancher, T., & Rowen, Z., H. (2007). Environmental effects of off-highway vehicles on Bureau of Land Management lands: A literature synthesis, annotated bibliographies, extensive bibliographies, and internet resources: U.S. Geological Survey. Open-File Report 2007-1353, 225p. Romeril, M. (1985). Tourism and the environment – Towards a symbiotic relationship. International Journal of Environmental Studies, 25: 215-218. Saarinen, J. (2016). Political ecologies and economies of tourism development in Kaokoland, North-West Namibia. In Mostafanezhad, M., Carr, A. & R. Norum (eds.) Political ecology of tourism: Communities, power and the environment, pp. 213–230. London: Routledge. Saarinen, J. (2018). What are wilderness areas for? Tourism and political ecologies of wilderness uses and management in the Anthropocene. Journal of Sustainable Tourism. doi: https://doi.org/10.1080/09669582.2018.1456543. Santarém, F., Campos, J., Pereira, P., Hamidou, D., Saarinen, J., & Brito, J.C. (2018). Using multivariate statistics to assess ecotourism potential of water-bodies: A case- study in Mauritania. Tourism Management 67, 34-46. Santarém, F., & Paiva, F. (2015). Conserving desert biodiversity through ecotourism. Tourism Management Perspectives 16, 176-178. 62 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Santarém, F., Pereira, P., Saarinen, J., & Brito, J.C. (2019). New method to identify and map flagship fleets for promoting conservation and ecotourism. Biological Conservation 229, 113-124. Serra, G. (2007). Ecotourism in Palmyra Desert, Syria. A feasibility study. Birdlife International. 88pp. United Nations Environmental Programme (UNEP). (2006a). Tourism and deserts: A practical guide to managing the social and environmental impacts in the desert recreation sector (eds. de Assis, H. R., & Manca M.). Accessed on 15 January 2018 at http://wedocs.unep.org/handle/20.500.11822/2229. United Nations Environmental Programme (UNEP). (2006b). Global deserts outlook (ed. E. Ezcurra). Accessed on 11 January 2018 at http://wedocs.unep.org/handle/20.500.11822/9581. United Nations Development Programme (UNDP). (2011). The forgotten billion: MDG Achievement in the drylands. Accessed on 21 December at http://www.undp.org/content/undp/en/home/librarypage/environment- energy/sustainable_land_management/the_forgotten_billionmdgachievementinthedr ylands.html. United Nations World Tourism Organization (UNWTO). (2018). Promotion of sustainable tourism, including ecotourism, for poverty eradication and environment protection. Accessed on 21 December 2018 at http://sdt.unwto.org/unga-sustainable-tourism- resolutions. United Nations World Tourism Organization (UNWTO)., & United Nations Development Programme (UNDP). (2017). Tourism and the Sustainable Development Goals – Journey to 2030. Accessed on 21 December of 2018 at http://www2.unwto.org/publication/tourism-and-sustainable-development-goals- journey-2030. United Nations Educational, Scientific and Cultural Organization (UNESCO). (2018). World Heritage List. Accessed on 23 December at https://whc.unesco.org/en/list/. Wagar, A. (1964). The carrying capacity of wild lands for recreation. Forest Science Monographs 7, 1–24. Waldron, A., Mooers, A. O., Miller, D. C., Nibbelink, N., Redding, D., Tyler, S. K., Timmons Roberts, J., &, Gittleman, J. L. (2013). Targeting global conservation funding to limit immediate biodiversity declines. PNAS 110, 12144-12148. Watson, J. E. M., Venter, O., Lee, J., Jones, K. R., Robinson, J. G., Possingham, H. P., & Allan, J. R. (2018). Protect the last of the wild. Nature 563, 27-30. Wearing, S. (2001). Exploring socio-cultural impacts on local communities. In Weaver, D. (Ed.) The Encyclopedia of Ecotourism. Wallingford, UK: CAB International. FCUP 63 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Weaver, D. B. (2001a). Ecotourism. Queensland, Australia: John Wiley & Sons Australia, Ltd. Weaver, D. B. (2001b). Deserts, and savannahs. In Weaver, D. B. (Ed.) The Encyclopedia of Ecotourism. Wallingford, UK: CABI Publishing. 64 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Chapter 2

Objectives and Thesis Structure FCUP 65 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

2.1. General objectives

The main objective of this Thesis is to understand how ecotourism could work as a sustainable solution for desert conservation and local livelihoods improvement in remote deserts subjected to political instability and exhibiting rare and threatened fauna. Based on species optimal aesthetical traits to marketing purposes and on assessments of ecotourism development potential and constrains at local (inland water-bodies of Mauritania) and sub-continental (18 Sahara-Sahel range countries) scales, the multi- scale approach develops research on conservation planning options for ecotourism development and management in desert ecosystems for achieving environmental conservation and sustainability. In particular, it is aimed to answer the following questions: 1. Which traits characterise a flagship species amongst candidate Sahara-Sahel vertebrates, which groups of flagship species could be used in marketing ecotourism and conservation campaigns, and where are located hotspots of these groups of flagships? 2. According to ecological and cultural characteristics, and to features that constrain the supply of cultural ecosystem services, which are the most suitable areas for ecotourism development in Mauritanian inland water-bodies (local scale) and in the 18 Sahara-Sahel range countries (sub-continental scale)? 3. Given the spatial variability of ecological, cultural and constraint features, and of the socio-economic indicators of Sahara-Sahel countries, which are the countries’ performance in preserving and promoting the cultural services that their ecosystems supply? 4. Which will be the impacts of climate change on the identified tourism hotspots and which places will be unreliable for human survivability?

2.2. Thesis structure

The Thesis is organized in six Chapters, followed by appendices, as follows:

In Chapter 1 it is briefly presented the history of ecotourism, the debate over its many definitions, the confusion with other types of tourism, market segmentation within ecotourism and the ecotourism research agenda in different biomes. It is also presented 66 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

an introduction to the field of ecotourism, its research and development in deserts, the potentials and pitfalls of deserts to develop ecotourism, the challenges deserts present to develop ecotourism, and the research that is needed to conduct related to desert ecotourism. The review of these topics is published as a Book Chapter: Article I: Santarém, F., Saarinen, J., Brito, J.C. (20). Desert Conservation and Management: Ecotourism. In: Goldstein MI, DellaSala DA (Eds.), Encyclopedia of the World's Biomes, vol. 2. Elsevier, pp. 259-273. doi: https://doi.org/10.1016/B978-0-12-409548-9.11827-5.

In Chapter 2 (this chapter) it is presented the main objectives of the Thesis and its structure.

In Chapter 3 it is presented a new method to identify and map flagship fleets for promoting conservation and ecotourism. The results are published in the following manuscript: Article II: Santarém, F., Pereira, P., Saarinen, J., Brito, J.C. (2019). New method to identify and map flagship fleets for promoting conservation and ecotourism. Biological Conservation, 229: 113-124. doi: https://doi.org/10.1016/j.biocon.2018.10.017.

In Chapter 4 it is presented spatial analyses of ecotourism potential at local and sub-continental scales. The results are published in the two following manuscripts: Article III: Santarém, F., Campos, J.C., Pereira, P., Hamidou, D., Saarinen, J., Brito, J.C. (2018). Using multivariate statistics to assess ecotourism potential of waterbodies: A case-study in Mauritania. Tourism Management, 67: 34-46. doi: https://doi.org/10.1016/j.tourman.2018.01.001. Article IV: Santarém, F., Saarinen, J., Brito, J.C. (2020). Mapping and analysing cultural ecosystem services in conflict areas. Ecological Indicators, 110: 105943. doi: https://doi.org/10.1016/j.ecolind.2019.105943.

In Chapter 5 it is presented a conservation planning applied to Sahara-Sahel ecoregions, considering the socio-economic performance of desert countries to preserve cultural ecosystem services, as well as a first assessment of climate change impacts on Sahara-Sahel tourism hotspots. The results are published in the following manuscripts: FCUP 67 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Article V: Santarém, F., Saarinen, J., Brito, J.C. (2021). Assessment and prioritization of cultural ecosystem services in the Sahara-Sahelian. Science of the Total Environment, 777: 146053. doi: https://doi.org/10.1016/j.scitotenv.2021.146053. Article VI: Santarém, F., Gonçalves, D. Saarinen, J., Brito, J.C. (under review). Vulnerability of desert tourism to climate change: lessons from the Sahara-Sahel.

In Chapter 6 it is discussed how flagship species, spatial analyses at different scales, conservation planning considering socio-economic indicators and preparation for climate change impacts help to build strategies for ecotourism development in the Earth’s largest warm region. Future perspectives for research on desert ecotourism are also presented, as well as the application of the principles developed in this thesis for sustainable development in desert ecosystems.

Appendices to the chapters, containing supplementary material important to understand specific methods and results, are presented after the main chapters. An additional appendix with other works published during the time of Thesis is presented at the very end. 68 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Chapter 3

Ecotourism and Conservation Marketing

Article II

Santarém, F., Pereira, P., Saarinen, J., Brito, J.C. (2019). “New method to identify and map flagship fleets for promoting conservation and ecotourism”. Biological Conservation, 229: 113-124. FCUP 69 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Article II: New method to identify and map flagship fleets for promoting conservation and ecotourism

Frederico Santarém1,2,3, Paulo Pereira1,2, Jarkko Saarinen3,4, José Carlos Brito1,2

1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 – Geography Research Unit, University of Oulu, Finland 4 – School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

Abstract

Evaluating flagship species and their potential for biological preservation and ecotourism development is a key issue for many audiences within the conservation and social fields. Despite several methods available to identify flagships, their application is often constrained in remote, poorly studied regions. Developments are needed in statistical and spatially-explicit approaches to assess species’ traits influencing flagship appealing, to identify flagship fleets, and to map the location of flagship hotspots. Here, we developed a new method to identify flagship species in regions with knowledge gaps, using a two-stage statistical approach (ordination and clustering algorithms) to assess variable’s contribution to appealing and to group species sharing similar characteristics into flagship fleets. We then mapped areas concentrating the highest richness of flagships. Unique morphologies and behaviours, conservation status, endemicity, body size and weight, and feeding habits were the traits contributing the most to the flagship appealing. Nine flagship fleets were identified, from which two were the most suitable for conservation marketing and ecotourism promotion campaigns in Sahara-Sahel: Fleet A comprising 36 large-bodied species (18 mammals, 18 ) and Fleet B including 70 small-bodied species (10 birds, six mammals, 54 reptiles). A total of 19 and 16 hotspots were identified for large-bodied and small-bodied flagships, respectively. The methodology was suitable to identify flagship species for conservation marketing and for developing ecotourism operations in the Sahara-Sahel, to independently assess which 70 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

species’ traits are relevant for flagship appealing, and to organise fleets for multispecies- based marketing campaigns. The framework is scalable and replicable worldwide.

Keywords

Charismatic species; cluster analysis; deserts; flagship fleets; flagship hotspots; principal component analysis. FCUP 71 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Introduction

Flagship species are defined as “a species used as the focus of a broader conservation marketing campaign based on its possession of one or more traits that appeal to the target audience” (Veríssimo et al., 2011), serving as symbols to stimulate conservation awareness and action, drawing financial support to protect habitats and other species (Simberloff, 1998). Despite several reinterpretations by academics (Barua, 2011), the latest definitions focus on the strategic, socio-economic, and marketing character of the concept (Veríssimo et al., 2011; Walpole and Leader-Williams, 2002). Flagship species act as a marketing (the process of planning and executing the value and distribution of products and services between organizations and target audiences) strategic tool to influence target audiences’ preferences and behaviours for the benefit of conservation efforts, by placing them at the core of the marketing process (see Wright et al., 2015 for a full review). Thus, they have been used in conservation awareness (Veríssimo et al., 2009), fundraising (Clucas et al., 2008), ecotourism (Walpole and Leader-Williams, 2002) and community-based conservation initiatives (Barua et al., 2011), in the protection of species and habitats (Smith et al., 2012), in policy making (Jepson and Barua, 2015), and in the development of pro-conservation behaviours (Smith and Sutton, 2008). Flagship fleets were recently developed as a tool to group several species into one single, more successful, flagship marketing campaign. Thus, they can raise the profile of more than one species and benefit a wider range of biodiversity (Barua et al., 2011; Veríssimo et al., 2014), while ensuring that multiple stakeholder groups’ preferences (e.g. of desert ecotourists; avitourists; ‘Big-Five’ tourists; academics; conservationists; international NGOs) are covered in flagship campaigns (Di Minin et al., 2013; Lindsey et al., 2007). Flagship fleets were proposed to answer concerns regarding fund diversion from less to high charismatic species (e.g. Andelman and Fagan, 2000; Joseph et al., 2011) and to lead behavioural change within and between different audiences (Root-Bernstein and Armesto, 2013). Pooling species belonging to different taxonomic groups into one single fleet can help targeting the desires of different audiences, thus maximizing the return-on-investment in multiple-species conservation efforts (Veríssimo et al., 2014). Many criteria for selecting flagship species have been proposed (Bowen-Jones and Entwistle 2002) but selection is mostly based in morphological, behavioural and cultural traits that are likely to be appealing to the planned target audience (Batt, 2009; Barua et al., 2011, 2012; Veríssimo et al., 2009, 2011). These typically include, amongst 72 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

others, body size, endemicity level or conservation status A key aspect in flagship choice is the extent to which it builds attitudinal, behavioural, financial or political support, without which their effectiveness as flagships may be compromised (Veríssimo et al., 2011). Thus, flagship species can be selected according to different methodologies, following social marketing, environmental economics, or conservation biology objectives (Veríssimo et al., 2011). In the past, they were selected mostly based on charisma, i.e. cultural value, rather than on objective principles (Home et al., 2009). Several conservation initiatives selected flagships for marketing campaigns based on preconceptions about the types of species favoured by the public, including potential tourists, conservationists, and academics. However, in conservation marketing, product familiarity is critical to consumer preference and any research on flagship preferences should take into account the knowledge level of the audience for properly identifying flagships (Garnett et al., 2018). To overcome this issue, choice experiment (CE) approaches have been developed to identify flagships for ecotourism promotion and conservation marketing campaigns (e.g. Di Minin et al., 2013; Veríssimo et al., 2009, 2013). These tools are derived from marketing theory and explore how particular attributes of a product are valued (Brown, 2010) and capture the heterogeneity of respondents preferences towards biodiversity attributes (Di Minin et al., 2013; Hausmann et al., 2016; Naidoo and Adamowicz, 2005). However, to be informative, CE relies on the principle that the public is well informed about the available species diversity in a given area to take a decision about attractiveness. Understanding people’s preferences towards every possible option from a large species pool may be an impractical task and even be unfeasible in remote regions with knowledge gaps about local biodiversity (Veríssimo et al., 2009). Photography-based studies have been used to address this shortcoming (e.g., Macdonald et al., 2015, 2017), but confounding effects of differing photograph colouration and angle have been raised (Batt, 2009), suggesting that flagship selection needs methodological developments that allow refining the identification of suitable species for different flagship campaigns in regions where the public is not well informed about the local biodiversity (Garnett et al., 2018). Furthermore, the contribution of species traits to flagship appealing in CE approaches is evaluated only after respondents have chosen their preferred species, which is only possible in well-studied regions. In remote regions, non-heuristic methods using multi-criteria in large datasets of species may allow evaluating the suitability of a species for the development of conservation marketing and ecotourism campaigns (Veríssimo et al., 2011). Non-heuristic methods, such as Principal Component Analysis (PCA), depend only on the dataset traits to assess variables contribution, whereas heuristic methods are affected by the subjective judgments of individuals that can lead to different, even FCUP 73 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel antagonist answers to the same question (Pathak et al., 2017). Unsupervised learning methods, such as Cluster Analyses, are used to group a set of objects (e.g. species) based on the observed values of several variables for each individual object (e.g. Batt, 2009), in a way that objects in the same group are more similar to each other than to those in other groups (Tryon, 1939). The potential of non-heuristic methods has been successfully explored in grouping water-bodies with similar traits for ecotourism development (e.g. Santarém et al., 2018) or in bioregionalization exercises based in environmental variation and species distribution (e.g. Brito et al., 2016). Non-heuristic methods have the potential to eliminate biases from analytical processes (Batt, 2009) and thus may help identifying flagship fleets efficiently. In developing nations, often rich in endemic biota but lacking common charismatic megafauna (large-sized vertebrates; Hausmann et al., 2016), international ecotourism can be highly beneficial for the national gross domestic product and local development (UNESCO, 2003; Twining-Ward et al., 2018; Weaver, 2001). This is the case of Sahara-Sahel range countries in Africa that are categorised as low human development (Brito et al., 2014) and are underfunded for poverty alleviation and biodiversity loss retention schemes (Durant et al., 2012; Waldron et al., 2013), display extensive remote areas, lack large populations of common African flagship species (e.g. elephant, lion) but are rich in endemic vertebrates (Brito et al., 2016). Ecotourism marketing campaigns have been set as regional conservation priorities (Brito et al., 2016; Hosni, 2000), but detailed knowledge about biodiversity levels in the Sahara-Sahel is still limited (Brito et al., 2014), as well as people’s understanding of the local potential flagship species. Effective methodologies are needed to identify flagship species in contexts of endemic-rich desert developing countries lacking common flagships and exhibiting biodiversity knowledge gaps (Santarém and Paiva, 2015). Mapping the location of flagship hotspots, i.e., particular areas concentrating exceptional flagship species richness suitable for conservation marketing and ecotourism promotion (adapted from Marchese, 2015), is in need of development. These areas are defined by one or more species-based metrics (e.g., species richness, number of species restricted to a particular area, or functional diversity within the ecosystem) in order to protect species supporting unique roles (Marchese, 2015). Despite the potential to indicate the location of suitable regions to allocate flagship-based conservation initiatives and to positively influence public behaviour through biodiversity, mapping the richness of flagship species or flagship fleets remains unexplored. Species abundance data is scarce in Sahara-Sahel (Brito et al., 2014, 2016) but mapping hotspots in such desert areas would help setting priorities for ecotourism development and conservation, 74 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

which is highly relevant for such poorly known areas and for minimizing local species extinction (Vale et al., 2015; Durant et al., 2014). Here, we propose a systematic method to identify flagship species in regions where the public is not well informed about the local species diversity and where CE would be impossible to perform to identify flagships, using Sahara-Sahel vertebrates as case-study. We first assess objectively which species traits drive flagship appealing, then assess which fleets of species can be used in flagship campaigns, and finally map flagship hotspots. Particularly, we want to answer the following questions: 1) which species’ traits contribute the most to explain species’ flagship appealing?; 2) how many flagship fleets can be distinguished according to their shared characteristics to flagship suitability?; 3) which flagship fleets display potential characteristics for conservation marketing and ecotourism promotion campaigns?; and 4) where are located Flagship Hotspots in the Sahara-Sahel? Specifically, we hypothesize that: 1) physical attributes, appearance, likelihood of extinction and endemicity are the most relevant traits for flagship selection in the Sahara-Sahel; 2) the most important variables will help shaping flagship fleets given the variabilities in those traits; 3) species displaying similar traits will tend to be clustered together and form different fleets suitable for distinct conservation marketing and ecotourism promotion campaigns; and 4) flagship hotspots will generally tend to be concentrated in local biodiversity hotspots and will tend to spatially overlap hotspots of total species richness.

Methods

Study area

Includes the Sahara and the Sahel (≈11,200,000 km2) ecoregions (Dinerstein et al., 2017; Fig. A1) and comprises 4072 grid cells of 0.5 degree resolution (WGS84 coordinate reference system).

Species list and trait data

The list of continental vertebrates of the Sahara-Sahel and their distribution polygons was retrieved from IUCN (2017), and updated according to Brito et al. (2018, 2016). It comprises 1126 species, including 52 , 188 reptiles, 584 birds, and 302 mammals (Table A1). For each species, we collected data on 13 variables related to distributional variation, morphophysiological and behavioural characteristics, FCUP 75 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel conservation status, and cultural representation (Table 2.1), based on relevant literature and other sources (see Text A1). The variables were:

1) Area of occupancy (AOC): the size of the geographical distribution of the species under analysis has been suggested as a key criterion for the selection of flagship species, as species with narrow ranges can reinforce allegiance with the region and influence people’s willingness to pay for conserving low distributed animals, while large ranged species help promoting global priority areas (Barua et al., 2011; Bowen-Jones and Entwistle 2002; Root-Bernstein and Arnesto, 2013; Veríssimo et al., 2013). Using a Geographical Information System (GIS; ESRI, 2012), we calculated the area of occupancy as the area within a species extent of occurrence (area contained within the minimum convex polygon encompassing the known species locations) that is occupied by that species (IUCN, 2017); 2) Body size (BSIZE) and 3) body weight (BWEIG): these traits have been extensively used in other flagship selection studies, as larger sizes and weights usually influence the easiness to observe species in the wild and animal attractiveness (Barua et al., 2012; Clucas et al., 2008; Ebner et al., 2016; Home et al., 2009; Macdonald et .al., 2015, 2017; Smith et al., 2012; Veríssimo et al., 2014). We quantified the maximum body size (total length, TL) and the maximum weight irrespectively of sexual dimorphism in these variables (sexual differences in measurements were not considered; we recorded only the longest body length and the largest weight value of the two sexes); 4) Morphologies (MORPH) and 5) behaviours (BEHAV): species exhibiting unique characteristics are highly valued by the public (Jepson and Barua 2015; Macdonald et al., 2015; Root-Bernstein and Armesto, 2013; Veríssimo et al., 2009). We accessed morphological - such as keeled scales and tuft of hair on ears to protect against sand - and behavioural adaptations - such as sand swimming and adapted feathers to transport water - to local desert environments, as a proxy for these uniqueness's; 6) Number of colours (NCOL) and 7) colour patterns (COLPAT): species with complex body colorations and recognizable colour patterns are considered more appealing to international audiences (Batt, 2009; Barua et al., 2012, 2011; Macdonald et al., 2015). We quantified the maximum number of colours and colour patterns, irrespectively of ontogenetic shifts and sexual dimorphism. For coloration, we considered 10 basic colour schemes (white, black, yellow, green, blue, orange, grey, red, brown, and purple). For body pattern, we considered main patterns - uniform, patches, spots, and stripes (longitudinal or transverse bars) - and then quantified the cumulative number of patterns in each species: three (patches + spots + stripes), two (e.g. spots + stripes), one of those, or none (see Fig. A2 for examples); 76 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

8) Conservation status (CS): likelihood of extinction of a species has been extensively explored within the flagship literature because threatened species urge the development of international conservation campaigns to attract conservation funds and ecotourists are seen as the mean to attract the money needed (Barua et al., 2011; Bowen-Jones and Entwistle, 2002; Clucas et al., 2008; Ebner et al., 2016; Macdonald et al., 2015, 2017). Species conservation statuses were based on IUCN Red List of Threatened Species (IUCN, 2017). The Scimitar-horned Oryx (Oryx dammah) was considered in this study as Extinct in the Wild, following IUCN categories, but the species has been recently reintroduced in Chad (Brito et al., 2018); 9) Endemicity (END): species with restricted distribution provide symbols of regional and national adherence and reflect strong local identity (Bowen-Jones and Entwistle 2002; Takahashi et al., 2012; Veríssimo et al., 2009, 2014). Species were categorised as endemic from the Sahara-Sahel or as non-endemic (if the distribution covered areas outside the Sahara-Sahel); 10) Daily (ACTIV) and 11) seasonal (SEAS) activities: the easiness to observe a species is influenced by its daily and seasonal activity patterns (Macdonald et al., 2015; Root-Bernstein and Armesto, 2013; Veríssimo et al., 2013, 2009). Species that are easy to spot in the wild are usually preferable as flagships in comparison to others less conspicuous (Reynolds and Braithwaite 2001). We accessed daily activity by considering the species as being diurnal, nocturnal or cathemeral (day and night activity). We assessed species seasonal activity by evaluating if the species is sedentary (constantly present in the study area) or seasonal (exhibiting migratory movements and thus not available all year round); 12) Feeding habits (FEED): different species feeding habits attract distinct types of audiences that feel linked to one or another diet type (Ebner et al., 2016). Usually, carnivores are preferred by people who are raising the profile of potential flagships (Clucas et al., 2008; Macdonald et al., 2015, 2017). The trophic levels we considered were: carnivore (meat-eating, including and fishes), herbivore (vegetation protein-eating), omnivore (vegetation, fruit, seed, grain, and/or nectar and animal protein-eating), frugivore (fruit, seed, grain and/or nectar-eating), and necrophage (carrion-eating); 13) Cultural, religious and medical uses (CRM): the cultural significance of a species is a major characteristic to be foreseen in a flagship. Relationships to local art, folklore and handicraft, or uses of venoms for medical, or religious representations are highlighted in flagship species (Barua et al., 2012; Bowen-Jones and Entwistle 2002; Takahashi et al., 2012). We evaluated where species have any of those cultural, religious and/or medical uses described above and treated this as a binomial variable (yes or no). FCUP 77 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table 2.1. Variables used to evaluate the 1126 Sahara-Sahel species to be used in flagship marketing campaigns, their code, description, type (num: numerical; cat: categorical), units, and data sources. Variables Code Description Type Source

Area of AOC Area within a species' extent num IUCN, 2017; occupancy of occurrence which is BirdLife occupied in Sahara-Sahel: International and Number of half-degree cells NatureServe, 2017

Body size BSIZE Maximum body size (total num AmphibiaWeb, length): cm 2017; Jones et al., 2009; Encyclopedia of Life, 2017; Myhrvold et al., 2015; Text A1

Body weight BWEIG Maximum weight: gr num AmphibiaWeb, 2017; Jones et al., 2009; Encyclopedia of Life, 2017; Myhrvold et al., 2015; Text A1

Morphology MORPH Unique morphological num AmphibiaWeb, adaptations to desert 2017; Encyclopedia environments: Number of of Life, 2017; Text A1; expert- knowledge

Behaviour BEHAV Unique behavioural num AmphibiaWeb, adaptations to desert 2017; Encyclopedia environments: Number of of Life, 2017; Text A1; expert- knowledge

Colour NCOL Colours in the body: Number num AmphibiaWeb, number of 2017; Encyclopedia of Life, 2017; Text A1 78 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Colour COLPAT Colour patterns in the body: cat AmphibiaWeb, pattern uniform; patches; spots; or 2017; Encyclopedia stripes (1); pa+sp or pa+st or of Life, 2017; Text sp+st (2); pa+sp+st (3) A1 Conservation CS Conservation status: NE; cat IUCN, 2017 DD; LC; NT; VU; EN; CR; EW

Endemic END Endemic in the study area: cat IUCN, 2017 Yes/No

Activity ACTIV Activity patterns: diurnal, cat Jones et al., 2009; nocturnal or cathemeral Text A1; expert- (active in both periods of the knowledge day)

Seasonality SEAS Annual activity of species cat Text A1 that influences the availability and easiness to observe the species: All- Year round; seasonal

Feeding FEED Feeding habits: cat Jones et al., 2009; Herbivorous; Frugivorous; Text A1 Omnivorous; Necrophagous; Carnivorous

Cultural use CRM Animal representations in cat Text A1; expert- cultural and religious art, knowledge and use of animal components for medicine: Yes/No

Statistical analyses

We applied two procedures to identify which traits are contributing the most to the appealing character of species for flagship marketing (see Text A2 for details). First, we performed a PCA for mixed (numerical and categorical) data (PCAmix), after standardised the numerical data, and then applied an orthogonal rotation to the PCA FCUP 79 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel loadings (Chavent et al., 2012). Contribution of variables is given by the squared loadings (correlation coefficients) of species traits with the first two rotated axes of the PCAmix. Squared loadings are the squared correlation for numerical variables and the correlation ratios for categorical variables with the rotated components, respectively (Chavent et al., 2012, 2017b). Flagship fleets, i.e. groups of species sharing similar characteristics for different flagship-based marketing campaigns, were identified using a model-based clustering approach (Scrucca et al., 2016). Fourteen multivariate normal mixture models with different parameterizations concerning the distribution, volume, shape and orientation of the covariance matrix of the multivariate data of the species were estimated by maximum likelihood using an expectation-maximization algorithm. The best model and the optimum number of clusters were chosen using the Bayesian Information Criterion (BIC) applied to the two first rotated axes of the PCAmix. The optimum number of clusters defined the number of different flagship fleets and for each cluster we matched a flagship fleet (Text A3). All analyses were performed in R version 3.4.3, using the functions PCAmix and PCArot of the package ‘PCAmixdata’ (Chavent et al., 2017a) for variables contribution, and ‘mclust’ of MClust package (Scrucca et al., 2016) for flagship fleets quantification. From the groups identified in the clustering analysis, we identified which ones are the most interesting flagship fleets for conservation marketing and ecotourism promotion campaigns. The identification of these fleets were based on the presence of traits (such as body size/weight, endemism, conservation status or unique adaptations; Clucas et al., 2008; Veríssimo et al., 2009, 2013) in the species that belong to each identified flagship fleet and that are commonly appreciated by different audiences (e.g. desert ecotourists; avitourists; academics; conservationists).

Mapping flagship hotspots

To identify and map flagship hotspots of the selected fleets for conservation marketing and ecotourism promotion campaigns, we first intersected species distribution polygons with grids of 0.5-degree resolution to generate matrices of species presence/absence by grid cell in a GIS (ESRI, 2012). A species was considered to occur in a cell if any portion of the species’ range overlapped the cell. Then, flagship hotspots for each of the fleets were obtained by summing the number of flagship species occurring in each grid cell. In order to identify the areas that maximize the likelihood of observing flagships belonging to multiple fleets, the distributions of retrieved fleets were overlapped to generate the combined flagship hotspots, representing 50% of the richness of each of 80 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

the two groups. Protected areas network (IUCN and UNEP-WCMC, 2018) was overlapped with the flagship hotspot map to identify gaps in the representation of hotspots in currently protected areas.

Results

Traits driving flagship appealing

Several variables contributed to the flagship appealing: unique morphologies and behaviours, conservation status and endemicity on the first dimension of the PCAmix, and body size, body weight, conservation status and feeding habits on the second dimension. The two first dimensions accounted for 17.23% of the variability (Table 2.2; Fig. 2.1).

Fig. 2.1. Squared loadings of species traits within the first and second axes (Dim 1 and Dim 2) of the rotated PCA with mixed data. The PCA is performed based on six numerical variables and seven categorical variables. See Table 2 for the correlation coefficients of the variables within the two rotated axes.

Identification of flagship fleets

The model with the highest BIC value (Text A3) was selected to identify the number of flagship fleets. The BIC increased with the number of clusters and reached the plateau at nine clusters, which was chosen as the optimum number of flagship fleets. Some groups were more cohesive than others (e.g. fleet H is very cohesive while fleet A FCUP 81 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel is heterogeneously dispersed along the two dimensions; Fig. 2.2), due to the weight of each species traits on grouping species into flagship fleets (Fig. 2.1). Each flagship fleet displays distinct characteristics potentially suitable for different flagship-based conservation and ecotourism initiatives and audiences’ preferences (Table A1).

Table 2.2. Correlation coefficients of species traits with the first two axes (Dimension 1 and Dimension 2; highly correlated values are in bold) of the Principal Component Analysis with mixed data (PCAmix), the eigenvalues and the percentage of explained variation of these two axes. See Fig. 2.1 for the squared loadings of the species traits within the two first rotated axes.

Species Traits Dimension 1 Dimension 2 Area of occupancy 0.183 -0.034 Body size 0.027 0.841 Body weight -0.048 0.692 Morphology 0.746 0.026 Behaviour 0.721 0.045 Colour number 0.005 -0.163 Colour patterns 0.111 0.012 Conservation 0.461 0.341 Endemic 0.343 0.000 Activity 0.019 0.111 Seasonality 0.000 0.003 Feeding 0.108 0.376 Cultural Use 0.004 0.088 Eigenvalue 2.160 2.149 Cumulative variance (%) 8.64 17.23

82 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 2.2. Flagship fleets identified by the Bayesian Information Criterion (BIC) applied to the two first rotated axes (Dim 1 and Dim 2) of Principal Component Analysis with mixed data (PCAmix; see Text A2 for methodological details). Large- bodied flagships (brown points) and small-bodied flagships (dark blue points) were used to map flagship hotspots. Suitability of flagship fleets for conservation and ecotourism campaigns

From the nine flagship fleets identified, fleets A and B were the most suitable for conservation and ecotourism campaigns in Sahara-Sahel (Fig. 2.2). Group A (N= 36 species), hereafter designated as large-bodied flagships, included mostly large and heavy mammals (18 species; e.g. Addax nasomaculatus) and reptiles (18 species; e.g. nigriventris), approximately half of them are regional endemics and exhibited some unique desert adaptations, more than a half is threatened with extinction, and most are herbivorous (Table 2.3). Group B (N= 70 species), hereafter designated as small- bodied flagships, includes small and light birds (10 species; e.g. Passer luteus), mammals (six species; e.g. Ctenodactylus vali) and reptiles (54 species; e.g. Acanthodactylus aureus), most are regional endemics, exhibit several desert adaptations, are not threatened, and are carnivorous. In total, 7.8% of mammals, 36% of reptiles, and 1.7% of birds occurring in the Sahara-Sahel were identified as flagships, whereas no single was identified.

FCUP 83 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table 2.3. List of species within flagship fleets suitable for conservation and ecotourism flagship marketing campaigns in the Sahara-Sahel and indication of the class, taxa, common name, endemic status (END; Sahara or Sahel), IUCN conservation status (CS), and flagship hotspot (numbered by location) where they can be observed (see Fig. A1 for details on the areas considered). Species that are distributed outside the flagship hotspots are signalled (Outside). Species data follow IUCN (2017).

Class Taxa Common name END CS Flagship Hotspot Fleet A – Large-bodied flagships Mammalia Addax nasomaculatus Addax Sahara CR 6 Ammotragus lervia Barbary sheep Sahara VU 1; 6; 7; 9; 12; 13; 14; 16; 17; 19 Equus africanus African wild ass - CR 1; 20 Eudorcas rufifrons Red-fronted gazelle Sahel VU 2; 3; 4; 8; 10; 11 Gazella cuvieri Cuvier's gazelle Sahara EN 12; 13; 15 Gazella Dorcas Dorcas gazelle Sahara VU 1; 6; 7; 8; 9; 10; 11; 12; 13; 14; 15; 16; 17; 19; 20 Gazella leptoceros Slender-horned gazelle Sahara EN 20 Giraffa Camelopardalis Giraffe - LC 3; 4; 5 Hippopotamus amphibious Common hippopotamus - VU 2 Kobus megaceros Nile lechwe - EN 2 Loxodonta Africana African elephant - VU 1; 2; 3; 4; 5; 8 Nanger dama Dama gazelle Sahara CR 6; 7 Nanger soemmerringii Grant's gazelle - VU 1 Oryx beisa Fringe-eared oryx - NT 1 Oryx dammah Scimitar-horned oryx Sahara EW 19 Panthera leo African lion - VU 2; 3; 4; 5 84 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Syncerus caffer African buffalo - LC 2; 3 Trichechus senegalensis African manatee - VU 8; 10 Reptilia Acanthodactylus spinicauda Doumergue's fringe-fingered Sahara CR 13 lizard Crocodylus niloticus Nile crocodile - LC 1; 2; 20; 21 Crocodylus suchus West-African crocodile - NE 2; 3; 4; 5; 8; 10; 11; 19 Naja nubiae Nubian spitting cobra Sahel NE 1; 7; 19; 20 Pseudocerastes fieldi Field's horned viper - LC Outside Pseudocerastes persicus Perisan horned viper - LC Outside Python sebae Royal python - NE 1; 2; 3; 4; 5; 8; 10; 11 Testudo kleinmanni Egyptian tortoise Sahara CR 20 Uromastyx acanthinura Schmidt’s spiny-tailed lizard Sahara NE 13; 15 Uromastyx aegyptia Egyptian spiny–tailed lizard - VU 20 Uromastyx alfredschmidti Schmidt’s mastigure Sahara NT 16 Uromastyx dispar Sudan mastigure Sahara NE 2; 9; 11; 12; 14; 16; 17; 21 Uromastyx geyri Geyr’s spiny-tailed lizard Sahara NE 7; 9; 14 Uromastyx nigriventris Moroccan Spiny-tailed Lizard Sahara NE 12; 13; 15 Uromastyx occidentalis Giant spiny-tailed lizard Sahara NE 12 Uromastyx ocellata Ocellated Spinytail - LC 1; 20; 21 Uromastyx ornata Ornate mastigure - LC Outside Varanus griseus Desert monitor - NE 1; 6; 7; 9; 11; 12; 13; 14; 15; 16; 17; 20; 21 FCUP 85 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fleet B – small-bodied flagships Aves Passer cordofanicus Kordofan Sparrow Sahel LC Outside Passer luteus Sudan Golden Sparrow Sahel LC 5; 7; 8; 9; 10; 11; 12 Pterocles alchata Pin-tailed sandgrouse - LC 12; 13; 15 Pterocles coronatus Crowned sandgrouse - LC 1; 7; 9; 11; 12; 13; 14; 15; 16; 18; 20 Pterocles exustus Chestnut-bellied sandgrouse - LC 5; 7; 8; 9; 10; 11; 20 Pterocles gutturalis Yellow-throated sandgrouse - LC Outside Pterocles lichtensteinii Lichtenstein's sandgrouse - LC 1; 7; 9; 12; 13; 14; 16; 20 Pterocles orientalis Black-bellied sandgrouse - LC 12; 13; 15 Pterocles quadricinctus Four-banded sandgrouse - LC 5; 8; 10 Pterocles senegallus Spotted sandgrouse - LC 1; 9; 11; 12; 13; 14; 15; 16; 18; 20; 21 Mammalia Crocidura tarfayensis Tarfaya Shrew Sahara DD 12 Ctenodactylus vali Val's gundi Sahara DD 13 Ictonyx libyca Libyan Striped Weasel Sahara LC 5: 7; 8; 10; 11; 12; 13; 15; 16; 18; 20 Jaculus jaculus Lesser Egyptian jerboa - LC 5; 7; 8; 10; 11; 12; 13; 14; 15; 16; 18; 20 Jaculus orientalis Greater Egyptian jerboa - LC 15; 20 Vulpes zerda Fennec fox Sahara LC 5; 7; 8; 9; 10; 11; 12; 13; 14; 15; 16; 17; 18; 20; 21 86 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Reptilia Acanthodactylus aegyptius Egyptian fringe-fingered lizard Sahara NE 18; 20 Acanthodactylus aureus Golden fringe-fingered lizard Sahara NE 12 Acanthodactylus boskianus Bosc’s fringe-toed lizard - NE 1; 5; 7; 8; 9; 10; 11; 12; 13; 14; 15; 16; 18; 20; 21 Acanthodactylus dumerili Duméril's fringe-fingered lizard Sahara NE 8; 9; 10; 11; 12; 13; 14; 15 Acanthodactylus longipes Long fringe-fingered lizard Sahara NE 7; 11; 12; 13; 14; 15; 16; 18; 20 Acanthodactylus Arnold's fringe-fingered lizard - LC Outside opheodurus Acanthodactylus scutellatus Nidua fringe-fingered lizard - NE 7; 9; 11; 13; 14; 15; 16; 18; 20; 21 Acanthodactylus taghitensis Taghit's fringe-toed lizard Sahara DD 12 Agama boueti Bouet's agama Sahel LC 5; 7; 8; 9; 10; 11; 12 Agama boulengeri Boulenger's agama Sahara LC 11 Agama spinosa Spiny Agama - LC 1; 20 Agama tassiliensis Tassili agama Sahara LC 7; 9; 14; 16 Cerastes cerastes Desert horned viper - NE 1; 7; 9; 11; 12; 13; 14; 15; 16; 18; 20 Cerastes vipera Sahara sand viper - LC 7; 8; 9; 10; 11; 12; 13; 14; 15; 16; 18; 20 Chalcides boulengeri Boulenger's feylinia Sahara NE 12; 13; 14; 15; 16; 18 Chalcides delislei De l'Isle's wedge-snouted skink Sahel LC 1; 5; 7; 8; 9; 10; 11; 12; 21 Chalcides humilis Ragazzi's bronze skink Sahara NE 1; 20 Chalcides sepsoides Wedge-snouted skink- Sahara LC 18; 20 FCUP 87 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Chalcides sphenopsiformis Duméril's wedge-snouted skink Sahara LC 12 Dasypeltis sahelensis Sahel egg eater Sahel NE 5; 7; 8; 10; 11; 12 Echis coloratus Palestine saw-scaled viper - NE 1; 20 Echis pyramidum Egyptian saw-scaled viper - LC 5; 7; 8; 9; 10; 11; 12; 14; 21 Eryx jaculus Javelin Sand Boa - NE 18; 20 Leptotyphlops algeriensis Beaked thread-snake Sahara NE 7; 11; 12; 16 Leptotyphlops boueti Bouet’s worm snake Sahel NE 8; 10 Leptotyphlops cairi Two-coloured blind snake Sahara NE 20; 21 Leptotyphlops Beaked blind snake - NE 18; 20 macrorhynchus Leptotyphlops nursii Nurse's blind snake - NE Outside Mauremys leprosa Mediterranean Turtle - NE 12; 13; 15 Mesalina rubropunctata Red-spotted lizard Sahara NE 7; 9; 11; 12; 13; 14; 18; 20 Philochortus lhotei Lhote's shield-backed ground Sahel NE 7; 14 lizard Pristurus adrarensis Adrar semaphore Sahara DD 11 Pseudotrapelus sinaitus Sinai agama - NE 20 Ptyodactylus guttatus Sinai fan-fingered gecko - NE 20 Ptyodactylus hasselquistii Yellow fan-fingered gecko - NE 20 Ptyodactylus oudrii Oudri's fan-footed gecko - LC 12; 13 Ptyodactylus ragazzi Ragazzi’s fan-footed gecko - NE 5; 7; 8; 9; 10; 11; 14; 15; 21 Ptyodactylus siphonorhina Sinai fan-fingered gecko Sahara NE 5; 8; 9; 10; 11; 21 88 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Scincopus fasciatus Peters' banded skink Sahel DD 5; 7; 8; 9; 10; 12; 13; 15; 21 Scincus albifasciatus Senegal Sandfish Sahara NE 8; 10; 11; 12; 13 Scincus scincus Sandfish skink - NE 9; 13; 14; 15; 16; 18; 20; 21 Stellagama stellio Starred Agama - LC 20 Stenodactylus petri Egyptian sand gecko Sahara NE 1; 7; 9; 11; 12; 13; 14; 15; 16; 18; 20 Stenodactylus stenodactylus Elegant Gecko Sahara NE 1; 7; 9, 11; 12; 13; 14; 15; 16; 18; 20 Trapelus boehmei Desert agama Sahara LC 11; 12; 13 Trapelus mutabilis Desert agama Sahara NE 7; 9; 13; 14; 15; 16; 17, 18; 20 Trapelus pallidus Pallid agama - NE 20 Trapelus schmitzi Schmitz’ agama Sahara DD 14; 16 Trapelus tournevillei Sahara agama Sahara LC 13; 15 Tropiocolotes algericus Algerian sand gecko Sahara NE 12; 13; 15 Tropiocolotes bisharicus Bishari pigmy gecko Sahara NE 1 Tropiocolotes steudeneri Algerian sand gecko Sahara NE 1; 7; 14; 16; 18; 20; 21 Typhlops etheridgei Mauritanian Blind Snake Sahara DD 11 Typhlops vermicularis European blind snake - NE 20

FCUP 89 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Flagship Hotspots

A total of 19 hotspots were identified for large-bodied flagships (regions number 1, 2, 3, 4, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 19, 20, and 21 in Fig. A1) and 16 hotspots for small-bodied flagships (regions number 1, 5, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 20, and 21 in Fig. A1). The combined map for large- and small-bodied flagships indicates the presence of flagship hotspots for both fleets combined in the mountains of Aïr, Hoggar, Tassili n’Ajjer and Mauritania, the Grand Erg Occidental, the Complex of Chotts, and the Nile River valley, and in sections of the Eastern Mountains, of the rivers White Nile, Niger and Senegal, and of the Western Corridor (see Fig. A1 for details). In both groups, flagship hotspots were located in particular Sahara-Sahel regions, such as mountains and waterbodies and are poorly overlapped by the current protected areas network (Fig. 2.3; Fig. A1).

90 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 2.3. Flagship Hotspots of large-bodied flagship fleets (above), small-bodied flagship fleets (centre), and both fleets combined (bottom) in the Sahara-Sahel. Areas with higher flagship richness are depicted in orange (above and centre) and dark red (bottom). The extent of Flagship Hotspots covered by current protected areas network (UNEP-WCMC and IUCN, 2018) is indicated by the Protected Areas polygons in the combined map. See Fig. A1 for the names of the areas.

FCUP 91 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Discussion

Methodological improvements from previous approaches

The two-step statistical approach here used, allowed the identification of flagship species and fleets in a systematic way. By assessing statistically which species traits contribute to flagship appealing, we answered researchers’ call for effective flagship evaluation methods that avoid biases inherent to people’s responses to questionnaires (Home et al., 2009; Veríssimo et al., 2011). Our approach is independent from questionnaire-based assessments, which would be impractical in the Sahara-Sahel anyway, as many species are unknown to the public but may display currently hidden characteristics suitable for flagships (Smith et al., 2012). The use of PCA with mixed data allowed assessing the contribution of numerical and categorical variables and improved the evaluation of traits contributing to flagship appeal. The use of clustering algorithms allowed identifying several flagship species sharing similar characteristics (Batt, 2009) that can be used in multispecies-target campaigns, extending the benefits of future marketing campaigns to a wider range of biodiversity (Veríssimo et al., 2014). The identification of flagship fleets considered all taxonomic groups together. The alternative strategy, analysing each taxonomic group individually, would likely inflate the importance of traits inside each taxonomic group that do not necessarily represent attractiveness in the real world. In the Sahara-Sahel, for instance, the area of occupancy for amphibians is a pointless variable due to the general desert environment of the region where few waterbodies are available to find them. -based analyses may also deflate the importance of traits that are commonly selected as attractive in real world situations identifying flagships (see Section 4.3 below). Pooling taxonomic groups onto the analysis allowed the statistics finding which are the most variable traits and how many fleets can be defined based on their intrinsic traits, thus not biasing (inflating or deflating) traits by taxonomic group and answering the need to identify flagship species objectively and independently of the taxonomic group to which they belong to (Santarém and Paiva, 2015; Veríssimo et al., 2014). Moreover, by pooling all taxa indiscriminately of their taxonomic group, we increased the chances of targeting higher number of different audiences at the same time, which may benefit a wider range of biodiversity (Di Minin et al., 2013; Hausmann et al., 2016). The use of GIS approach allowed mapping the location of flagship hotspots, which may guide the implementation of flagship-based conservation initiatives. The methodology can be extrapolated to other biomes and the spatial approach here used

92 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

constitutes a novel method that should be further explored in ecotourism-based research (Santarém et al., 2018).

Potential methodological improvements for future works

The method here used to identify flagship hotspots contains some potential caveats. Including variables associated with the perceptions of local people about flagships may improve flagship identification. For instance, species unlinked with human- wildlife conflicts may be acknowledged as important flagships, while crop-damaging species or that kill humans may be difficult to promote as flagships with positive symbolic value (Barua et al., 2010). The ecological and economic roles (e.g. pollination and food provision, respectively), and population size (as a proxy for rareness) may also catalyse conservation actions efficiently (Barua et al., 2011; Ebner et al., 2016; Veríssimo et al., 2009). Supervised machine-learning methods (Kotsiantis, 2007) can be used as alternative to the clustering method here applied to identify flagships. Based on a training set of successful flagship species together with species with low public appeal, models are trained to classify any species as flagship or not. Still, knowledge about the species under analysis must be available for effective method application (Smith et al., 2012). Given the general paucity in the availability of accurate biodiversity distribution data in the Sahara-Sahel (Brito et al., 2014), the mapping of flagship hotspots was based in IUCN species distribution polygons. IUCN polygons depict the species full extent of occurrence and naturally include non-occurrence areas within the range, which probably inflated the rate of false presences (e.g., Graham and Hijmans, 2006) and introduced omission and commission errors (Macdonald et al., 2017). In contrast, the most remote areas of the Sahara-Sahel or the areas subjected to long-term local conflicts are likely under sampled (Brito et al., 2016), which probably impacted the accuracy of the final mapped results. For instance, the combined map also demonstrates a completely absence of flagship hotspots in Libya and Chad and a poor representation in Sudan (Fig. 2.3). Additional research in these regions may dictate a different pattern, particularly in the data deficient Sudanese mountainous areas (Siddig, 2014), which may follow the general trend of flagship hotspots to be located in mountains (see Fig. 2.3 and Fig. A1 together). Although these constraints may cause some bias for the identified flagship hotspots due to deficient data, these effects were diluted when applying a coarse spatial resolution (pixel size of 0.5º). Still, future developments should target the use of accurate distribution data, for instance by deriving species ranges from ecological niche-based models. Additionally, the spatial and temporal dynamics of many flagships, such as FCUP 93 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel migrating birds or hibernating reptiles, should be contemplated in future developments, as these influence the likelihood of species observation and consequently the location of flagship hotspots.

Factors relevant for assessing flagship species in deserts

When identifying flagship species, there are species traits that are highly related. For instance, endemic species tend to display unique adaptations to deserts and large species are generally heavy (see correlations in axes XX and YY in Fig. 2.1, respectively). These relationships helped to systematically build inferences on what characteristics are relevant for flagship appealing when applying statistical methods, as we did here. Adaptations to deserts were relevant features in defining flagships in Sahara- Sahel (Fig. 2.1, Table 2.2), a finding similar to what was found in a study developed in Seychelles (Veríssimo et al., 2009), where species with unusual characteristics were selected by respondents. International tourists and ecotourism operators often prefer species with unique adaptations to the local environment as they turn out to be regional mottos where they live and hence can raise funds more efficiently than others without particular adaptations (Root-Bernstein and Arnesto, 2013; Veríssimo et al., 2009). Similarly, endemism stands out as an important feature for desert flagship species. Endemism was selected by local communities as an important feature for potential Brazilian flagships (Veríssimo et al., 2014), although tourists did not find it an important trait for Indian flagships (Takahashi et al., 2012). Rarity is also an attractive trait to the public that is willing to pay additional fees for ecotourism and conservation projects targeting endemic species (Veríssimo et al., 2009). This is an encouraging result for endemic-rich areas, such as Sahara-Sahel, composed by several developing countries where flagship-based ecotourism can have a positive role (Brito et al., 2016, 2014). Body size and weight helped defining flagship species in Sahara-Sahel. Fascination with large animals is widely reported as a key-element in defining flagship appeal (Macdonald et al., 2015, 2017) and large and heavy animals are also preferred by international NGOs when implementing flagship-based conservation programmes (Clucas et al., 2008). However, even small-sized flagships are acknowledged by specific target groups of tourists (Ebner et al., 2016; Macdonald et al., 2017), which is relevant to the heterogeneous pattern we found between the identified flagship fleets in the Sahara-Sahel (fleet A – large-bodied flagships and fleet B – small-bodied flagships; Fig. 2.2, Table A1). Hence, even the smallest animals can be used in flagship marketing

94 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

campaigns if they exhibit traits that make them appealing flagships (Smith et al., 2012). For instance, the small-bodied flagships fleet is composed by species that display many unique adaptations to deserts, such as the reptiles of genera Acanthodactylus, Cerastes, and Scincus, which may be preferred by specialized audiences in deserts (e.g. desert ecotourists and conservationists) and targeted for specialized desert marketing campaigns (Santarém et al., 2018). Likelihood of extinction was also found to be highly associated to species with flagship characteristics in the Sahara-Sahel. This kind of ‘last chance to see’ tourism is typical for various ecotourism destinations with rapidly changing habitats (Lemelin et al., 2012) and as a phenomenon it is supported by many studies (e.g. Batt, 2009; Clucas et al., 2008; Ebner et al., 2016; Macdonald et al., 2015), which contrasts with the unimportance found in some works (e.g. Macdonald et al., 2017 and Smith et al., 2012). Notwithstanding, non-threatened species may still be suitable for flagship marketing campaigns, as common species may be promptly chosen as regional ambassadors by local communities when compared to infrequently encountered rare species (Takahashi et al., 2012). This has implications for Sahara-Sahel species, as many of them are not threatened or have not been yet evaluated (Brito et al., 2016).

Factors potentially relevant for defining flagship species in other biomes/scales

Several species traits were identified as less important in defining flagship species in the Sahara-Sahel but may be relevant in other biomes (e.g. tropical jungles) or other scales of analysis (e.g. local, national). Although area of occupancy was one of the least important variables in this study, a pattern consistent with findings of other works, still, it is widely accepted that species with narrow ranges can reinforce local allegiance and influence people’s willingness to pay for conserving low distributed animals, while large-ranged species help promoting global priority regions (Barua et al., 2011; Bowen-Jones and Entwistle, 2002; Root-Bernstein and Armesto, 2013; Veríssimo et al., 2013, 2009). Thus, distribution patterns should be contextualized when identifying flagship species in different biome/scale contexts, as conservation problems may vary accordingly. Despite we were unable to find a grouping pattern by diet, carnivores are generally preferred by international ecotourists and wider audiences, as people not facing wildlife damages generally revere them (Clucas et al., 2008; Macdonald et al., 2017). However, other feeding habits may be preferred by specific audiences and thus this needs to be further contextualized in order to consider different preferences. FCUP 95 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Body patterns were not important for many people from different regions of the world in a global assessment (Macdonald et al., 2015). However, striking colour patterns have been found as an important attribute of flagship appealing in other taxonomic groups, namely invertebrates (Barua et al., 2012). Unusual colour patterns may attract specific audiences as well, and thus deserve to be contextualized, as people from different context may differ in their preferences towards coloration and body patterns (Macdonald et al., 2015). Visibility traits followed similar results in bird flagships assessments (Veríssimo et al., 2009). This contrasts with the recommendations to use such feature in wildlife tourism, where diurnal activity pattern and predictable activity are considered influential to observers’ preferences (Reynolds and Braithwaite, 2001). However, some specialized tourist segments may prefer rare and difficultly observable species in Africa (Di Minin et al., 2013), which opens an opportunity to broadening flagship marketing campaigns in the Sahara-Sahel. It has been claimed that utilitarian values are not important to general tourists (Veríssimo et al., 2009), whereas local people put high value in the usefulness of species (Takahashi et al., 2012; Barua et al., 2012). International knowledge of species within the Sahara-Sahel is practically null (Brito et al., 2014), but there is an opportunity to explore this trait further once more scientific knowledge is raised. The use of poisons from desert species for medical research and cultural believes, and locally religious representations (e.g. old Animal-Gods in Egypt) can raise the profile of some neglected species, thus contributing to attract funds to its conservation through ecotourism initiatives. The criteria here used serve as a set of guidelines for what attributes should be used to identify flagship fleets that resonate with different audiences. Both tourists and locals need to be considered when assessing such features (Macdonald et al., 2017; Smith et al. 2009). Considering both audiences motivate them to support flagship conservation projects, with implications to broader biodiversity (Barua et al., 2012; Takahashi et al., 2012).

Sahara-Sahel flagship fleets

The variability contained in the most important variables helped shaping fleets. The cluster algorithm grouped several species sharing similar characteristics with flagship appealing into nine flagships fleets, which potentially allows promoting many Sahara-Sahel species in multi-flagships campaigns, highlighting the benefits of using flagship fleets to preserve more than just only the most popular species (Veríssimo et

96 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

al., 2014). From the fleets here identified, large- and small-bodied flagships, retaining many endemic species (e.g. Addax nasomaculatus in fleet A and Acanthodactylus aureus in fleet B; Table 2.3; Table A1), play a key role in the development of strategic marketing campaigns in Sahara-Sahel. Despite the shocking on-going extinction of Sahara-Sahel species (from which several of them were identified as flagships), the world is neglecting reversal conservation measures that would prevent their collapse (Durant et al., 2014; Brito et al., 2018). Hence, these two flagship fleets in particular can be used by local governments to raise the profile of currently overlooked species and to attract international conservation donors, as even less popular species might raise funds in special occasions (Hausmman et al., 2016; Macdonald et al., 2017; Veríssimo et al., 2017). Other ecotourism segments can be attracted as well, which may diversify and enhance tourism operations targeting alternative flagship species (Lindsey et al., 2007; Walpole and Leader-Williams, 2002). Other identified fleets may be suitable for particular initiatives, for instance groups H and I are constituted by several bird species, thus being optimal for specific birdwatching and NGO’s - BirdLife International - conservation campaigns, or for other segments of the society not specialized in desert biodiversity (Di Minin et al., 2013; Root- Bernstein and Armesto, 2013). Birds were poorly represented in the two most suitable fleets for flagship campaigns (fleets A and B) because there are few endemic birds in the Sahara-Sahel with adaptations to desert conditions (Appendix F). However, they might be strong flagship species in other regions rich in endemic birds (e.g., in tropical islands; Veríssimo et al., 2009). No single amphibian was identified as suitable flagships for conservation marketing and ecotourism promotion campaigns (fleets A and B), as only two of them are endemic to the region (Appendix F) and none displays adaptations to local conditions (Brito et al., 2016). Amphibians were investigated as candidate flagships in other works (e.g., Bride et al., 2008), presenting restricted opportunities for generalist audiences not primarily interested in observing regional desert endemic species and for amphibian-enthusiasts that wish to observe more species of this taxonomic group in the Sahara-Sahel. However, amphibians may be strong flagships for conservation and ecotourism stakeholders in other regions of the world, where amphibian richness is higher than in arid regions and where amphibian species display several adaptations to local conditions (e.g., in India; Kanagavel et al., 2017).

Flagship hotspots in the Sahara-Sahel

This was the first study mapping flagship hotspots that may attract ecotourists to these regions and that are able to fund conservation programmes targeting flagship FCUP 97 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel species (Santarém and Paiva, 2015; Walpole and Leader-Williams, 2002). Overall, hotspots for large-bodied flagships were mostly fragmented, while hotspots for small- bodied flagship tend to be spatially continuous. This pattern reflects the high regional fragmentation levels of large animal populations (Brito et al., 2014, 2018; Durant et al., 2014) and highlights the need for designing corridors connecting these fragmented patches to protect large-bodied animals (Brito et al., 2016). Although total species richness is larger along the Sahel (Brito et al., 2016), the combined map of hotspots for both flagship fleets suggests that the central Sahara accumulates the highest flagship richness (Fig. 2.3) and that flagship-based observation initiatives probably should be prioritized in this region. Despite the prediction that the Sahel should offer more animal encounters than the Sahara due to local total species richness, it is here demonstrated that the Saharan mountains and waterbodies maximize flagship richness, and that these places may be optimal for flagship-targeted marketing campaigns. The location of flagship hotspots is broadly coincident with priority conservation areas identified in Sahara-Sahel (Brito et al., 2016). Still, more than half of the flagship hotspots are not currently protected (Fig. 2.3), which urges the development of protecting schemes towards this unique desert biodiversity, as species occurring in areas of high conservation priority display added value for their conservation marketing (Macdonald et al., 2017). Flagship hotspots in Mauritania, Mali, Niger, and Egypt are particularly poorly protected. By identifying the areas concentrating more flagships, we complement the urgency of preserving these regions and provide local governments with an additional tool to raise funds through flagship species marketing campaigns (Santarém et al., 2018). As most of the species of large- and small-bodied flagship fleets correspond to the most threatened species in the Sahara-Sahel, ecological corridors and transboundary mega conservation areas should be prioritised if these species are to be saved (Brito et al., 2016).

Conclusions

Evaluating flagship species potential for biological conservation and ecotourism development is an increasing important focus area in many parts of the world, and flagship marketing remains a key fundraising tool for international agencies (e.g., IUCN and United Nations) and NGOs, local governments, and the scientific community. The approach developed here is scalable and replicable worldwide, and the used criteria serve as a set of guidelines for what attributes could and should be used to identify flagship fleets that resonate with different audiences. The methodology has implications

98 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

for the preservation of several species, as it allows the identification of flagship fleets for conservation marketing and ecotourism promotion in a systematic way. By doing so, flagship-rich regions (flagship hotspots) located in low-income countries can attract international donors able to fund conservation campaigns, hence benefiting even the less popular and charismatic species.

References

AmphibiaWeb. (2017). AmphibiaWeb. Available at https://amphibiaweb.org/ (accessed 3 March 2017). Andelman, S. J., & Fagan, W. F. (2000). Umbrellas and flagships: Efficient conservation surrogates or expensive mistakes? Proceedings of the National Academy of Sciences, 97 (11), 5954–5959. doi: 10.1073/pnas.100126797. Barua, M. (2011). Mobilizing metaphors: the popular use of keystone, flagship and umbrella species concepts. Biodiversity and Conservation, 20, 1427-1440. doi: 10.1007/s10531-011-0035-y. Barua, M., Gurdak., D. J., Ahmed, R. A., & Tamuly, J. (2012). Selecting flagship species for invertebrate conservation. Biodiveristy and Conservation, 21, 1457-1476. doi: 10.1007/s10531-012-0257-7. Barua, M., Root-Bernstein, M., Ladle, R. J., & Jepson, P. (2011). Defining flagship uses is critical for flagship selection: A critique of the IUCN Climate Change Flagship Fleet. Ambio, 40 (4), 431–435. doi: 10.1007/s13280-010-0116-2. Barua, M., Tamuly, J., & Ahmed, R. A. (2010). Mutiny or clear sailing? Examining the role of the Asian Elephant as a flagship species. Human Dimensions of Wildlife, 15, 145–160. doi: 10.1080/10871200903536176. Batt, S. (2009). Human attitudes towards animals in relation to species similarity to humans: a multivariate approach. BioscienceHorizons, 2 (2), 180-190. doi: 10.1093/biohorizons/hzp021. BirdLife International, & NatureServe. (2017). Bird species distribution maps of the world. Available at http://www.birdlife.org/datazone/info/spcdownload (accessed 11 November 2017). Bowen-Jones, E., & Entwistle, A. (2002). Identifying appropriate flagship species: The importance of culture and local contexts. Oryx, 36, 189–195. doi: 10.1017/S0030605302000261. Bride, I. G., Griffiths, R. A., Meléndez-Herrada, A., & McKay, J. E. (2008). Flying an amphibian flagship: conservation of the Axolotl Ambystoma mexicanum through FCUP 99 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

nature tourism at Lake Xochimilco, Mexico. International Zoo Yearbook, 42, 116- 124. doi: 10.1111/j.1748-1090.2008.00044.x. Brito, J. C., Durant, S. M., Pettorelli, N., Newby, J., Canney, S., Algadafi, W., … Carvalho, S. B. (2018). Armed conflicts and wildlife decline: Challenges and recommendations for effective conservation policy in the Sahara-Sahel. Conservation Letters, 11, 1– 13. doi: 10.1111/conl.12446. Brito, J. C., Godinho, R., Martínez-Freiría, F., Pleguezuelos, J. M., Rebelo, H., Santos, X., … Carranza, S. (2014). Unravelling biodiversity, evolution and threats to conservation in the Sahara-Sahel. Biological Reviews, 89, 215–231. doi: 10.1111/brv.12049. Brito, J. C., Tarroso, P., Vale, C. G., Martínez-Freiría, F., Boratyński, Z., Campos, J., … Carvalho, S. B. (2016). Conservation Biogeography of the Sahara-Sahel: Additional protected areas are needed to secure unique biodiversity. Diversity and Distributions, 22 (4), 371–384. doi: 10.1111/ddi.12416. Brown, S. (2010). Where the wild brands are: some thoughts on anthropomorphic marketing. Marketing Review, 10(3), 209‐224. Available at https://ssrn.com/abstract=2017467 (accessed 20 June 2018). Chavent, M., Kuentz-Simonet, V., Labenne, A., Liquet, B., & Saracco, J. (2017a). PCAmixdata: Multivariate Analysis of Mixed Data. R package version 3.1. Available at https://cran.r-project.org/package=PCAmixdata (accessed 13 March 2018). Chavent, M., Kuentz-Simonet, V., Labenne, A., & Saracco, J. (2017b). Multivariate Analysis of Mixed Data: The R Package PCAmixdata. Chavent, M., Kuentz-Simonet, V., & Saracco, J. (2012). Orthogonal rotation in PCAMIX. Advances in Data Analysis and Classification, 6, 131–146. doi: 10.1007/s11634- 012-0105-3. Clucas, B., Mchugh, K., & Caro, T. (2008). Flagship species on covers of US conservation and nature magazines. Biodiversity and Conservation, 17, 1517– 1528. doi: 0.1007/s10531-008-9361-0. Di Minin, E., Fraser, I., Slotow, R., & MacMillan, D. C. (2013). Understanding heterogeneous preference of tourists for big game species: implications for conservation and management. Animal Conservation, 16, 249-258. doi: 10.1111/j.1469-1795.2012.00595.x. Dinerstein, E., Olson, D., Joshi, A., Vynne, C., Burgess, N. D., Wikramanayake, E., … Kindt, R. (2017). An Ecoregion-Based Approach to Protecting Half the Terrestrial Realm. BioScience, 67 (6), 534–545. doi: 10.1093/biosci/bix014. Durant, S. M., Pettorelli, N., Bashir, S., Woodroffe, R., Wacher, T., Ornellas, P. de, … Baillie, J. (2012). Forgotten Biodiversity in Desert Ecosystems. Science, 336 (6087),

100 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

1379–1380. doi: 10.1126/science.336.6087.1379. Durant, S. M., Wacher, T., Bashir, S., Woodroffe, R., De Ornellas, P., Ransom, C., … Pettorelli, N. (2014). Fiddling in biodiversity hotspots while deserts burn? Collapse of the Sahara’s megafauna. Diversity and Distributions, 20, 114–122. doi: 10.1111/ddi.12157. Ebner, B. C., Morgan, D. L., Kerezsy, A., Hardie, S., Beatty, S. J., Seymour, J. E., … McAllister, R. R. J. (2016). Enhancing conservation of Australian freshwater ecosystems: Identification of freshwater flagship fishes and relevant target audiences. Fish and Fisheries, 17(4), 1134–1151. doi: 10.1111/faf.12161. Encyclopedia of Life. (2017). Encyclopedia of Life. Available at www.eol.org (accessed 14 August 2017). Environmental Systems Research Institute (ESRI). (2012). ArcGIS Release 10.1. CA: Redlands. Garnett, S. T., Ainsworth, G.B., & Zander, K.K. (2018). Are we choosing the right flagships? The bird species and traits Australians find most attractive. PLoS ONE, 13 (6): e0199253. doi: https://doi.org/10.1371/journal.pone.0199253. Graham, C. H., & Hijmans, R. J. (2006). A comparison of methods for mapping species ranges and species richness. Global Ecology and Biogeography, 15 (6), 578-587. doi: 10.1111/j.1466-822x.2006.00257.x. Hausmann, A., Slotow, R., Fraser, I., & Di Minin, E. (2016). Ecotourism marketing alternative to charismatic megafauna can also support biodiversity conservation. Animal Conservation, 20, 91-100. doi: 10.1111/acv.12292. Home, R., Keller, C., Nagel, P., Bauer, N., & Hunziker, M. (2009). Selection criteria for flagship species by conservation organizations. Environmental Conservation, 36 (2), 139–148. doi: 10.1017/S0376892909990051. Hosni, E. (2000). Strategy for sustainable tourism development in the Sahara. Paris: UNESCO. International Union for Conservation of Nature (IUCN). (2017). The IUCN Red List of Threatened Species, 2017-3. International Union for Conservation of Nature and Natural Resources. Available at http://www.iucnredlist.org/ (accessed 15 November 2017). IUCN & UNEP-WCMC. (2018). The World Database on Protected Areas (WDPA). Cambridge, UK: UNEP-WCMC. Available at http://www.protectedplanet.net (accessed 27 August 2018). Jepson, P., & Barua, M. (2015). A Theory of Flagship Species Action. Conservation and Society, 13, 95–104. doi: 10.4103/0972-4923.161228. Jones, K. E., Bielby, J., Cardillo, M., Fritz, S. A., O’Dell, J., Orme, C. D. L., … Purvis, A. FCUP 101 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

(2009). PanTHERIA: A species-level database of life history, ecology, and geography of extant and recently extinct mammals. Ecology, 90 (9), 2648. doi: 10.1890/08-1494.1. Joseph, L. N., Maloney, R. F., Watson, J. E. M., & Possingham, H. P. (2011). Securing nonflagship species from extinction. Conservation Letters, 4 (4), 324–325. doi: 10.1111/j.1755-263X.2011.00174.x. Kanagavel, A., Eluvathingal, L. M., Parvathy, S., Kotharambath, R., & Das., S. (2017). Potential flagship species for improving support and garnering attention towards amphibian conservation in the Western Ghats, India, in Das, A. (ed.), Diversity & Ecology of Amphibians of India, ENVIS Bulletin: Wildlife & Protected Areas, Volume 19. Dehradun. Wildlife Institute of India, pp. 271–292. Kotsiantis, S. B. (2007). Supervised machine learning: A review of classification techniques. Informatica, 31, 249-268. Lemelin, H., Dawson, J., & Stewart, E.J. (2012). Last-Chance Tourism: Adapting tourism Opportunities in a changing World. London: Routledge. Lindsey, P. A., Alexander, R., Mills, M. G. L., Romañach, S., & Woodroffe, R. (2007). Wildlife Viewing Preferences of Visitors to Protected Areas in South Africa: Implications for the Role of Ecotourism in Conservation. Journal of Ecotourism, 6, 19-33. doi: 10.2167/joe133.0. Macdonald, E. A., Burnham, D., Hinks, A. E., Dickman, A. J., Malhi, Y., & Macdonald, D. W. (2015). Conservation inequality and the charismatic cat: Felis felicis. Global Ecology and Conservation, 3, 851-866. doi: 10.1016/j.gecco.2015.04.006. Macdonald, E. A., Hinks, A. E., Weiss, D. J., Dickman, A. J., Burnham, D., Sandom, C. J., …, & Macdonald, D. W. (2017). Identifying ambassador species for conservation marketing. Global Ecology and Conservation, 12, 204-214. doi: 10.1016/j.gecco.2017.11.006. Marchese, C. (2015). Biodiversity hotspots: A shortcut for a more complicated concept. Global Ecology and Conservation, 3, 297-309. doi: 10.1016/j.gecco.2014.12.008. Myhrvold, N. P., Elita, B., Benjamin, C., Sivam, D., Freeman, D. L., & Ernest, S. K. M. (2015). An amniote life-history database to perform comparative analyses with birds, mammals, and reptiles. Ecology, 96 (11), 3109. doi: 10.1890/15-0846R.1. Naidoo, R. & Adamowicz, W. L. (2005). Biodiversity and nature-based tourism at forest reserves in . Environment and Development Economics, 10, 159-178. doi: 10.1017/S1355770X0400186X. Pathak, S., Ojha, C., Zevenbergen, C., & Garg, R. (2017). Ranking of storm water harvesting sites using heuristic and non-heuristic weighing approaches. Water, 9 (9), 710. doi: 10.3390/w9090710.

102 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Reynolds, P. C., & Braithwaite, D. (2001). Towards a conceptual framework for wildlife tourism. Tourism Management, 22, 31–42. doi: 10.1016/S0261-5177(00)00018-2. Root-Bernstein, M., & Armesto, J. (2013). Selection and implementation of a flagship fleet in a locally undervalued region of high endemicity. Ambio, 42 (6), 776-787. doi: 10.1007/s13280-013-0385-7. Santarém, F., Campos, J. C., Pereira, P., Hamidou, D., Saarinen, J., & Brito, J. C. (2018). Using multivariate statistics to assess ecotourism potential of water-bodies: A case- study in Mauritania. Tourism Management, 67, 34–46. doi: 10.1016/j.tourman.2018.01.001. Santarém, F., & Paiva, F. (2015). Conserving desert biodiversity through ecotourism. Tourism Management Perspectives, 16, 176-178. doi: 10.1016/j.tmp.2015.07.016. Scrucca, L., Fop, M., Murphy, T. B., & Raftery, A. E. (2016). mclust 5: Clustering, classification and density estimation using Gaussian finite mixture models. R Journal, 8, 289–317. Siddig, A. A. H. (2014). Biodiversity of Sudan: between the harsh conditions, political instability and civil wars. Biodiversity Journal, 5 (4), 545-555. Available at: http://www.biodiversityjournal.com/ (accessed 28 August 2018). Simberloff, D. (1998). Flagships, umbrellas, and keystones: Is single-species management passé in the landscape era? Biological Conservation, 83 (3), 247– 257. doi: 10.1016/S0006-3207(97)00081-5. Smith, A. M., & Sutton, S. G. (2008). The Role of a Flagship Species in the Formation of Conservation Intentions Conservation Intentions. Human Dimensions of Wildlife, 13 (2), 127–140. doi: 10.1080/10871200701883408. Smith, R. J., Veríssimo, D., Isaac, N. J. B., & Jones, K. E. (2012). Identifying Cinderella species: uncovering mammals with conservation flagship appeal. Conservation Letters, 5 (3), 205–212. doi: 10.1111/j.1755-263X.2012.00229.x. Smith, R. J., Veríssimo, D., Leader-Williams, N., Cowling, R. M., & Knight, A. T. (2009). Let the locals lead. Nature, 462, 280-281. doi: 10.1038/462280a. Takahashi, Y., Veríssimo, D., MacMillan, D. C., & Godbole, A. (2012). Stakeholder perceptions of potential flagship species for the Sacred Groves of the North Western Ghats, India. Human Dimensions of Wildlife, 17 (4), 257–269. doi: 10.1080/10871209.2012.675622. Tryon, R. C. (1939). Cluster analysis: Correlation Profile and Orthometric (Factor) Analysis for the Isolation of Unities in Mind and Personality. Edwards Brothers, Ann Arbor. Twining-Ward, L. D., Wendy, L., Elisson M. W., & Mukeshkumar, B. H. (2018). Supporting sustainable livelihoods through wildlife tourism (English). Tourism for FCUP 103 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Development knowledge series. Washington, D.C.: World Bank Group. Available at http://documents.worldbank.org/curated/en/494211519848647950/Supporting- sustainable-livelihoods-through-wildlife-tourism (accessed 28 August 2018). Vale, C., Pimm, S. L., & Brito, J. C. (2015). Overlooked Mountain Rock Pools in Deserts Are Critical Local Hotspots of Biodiversity. PLoS ONE, 10 (2), e0118367. doi: 10.1371/journal.pone.0118367. Veríssimo, D., Fraser, I., Girão, W., Campos, A. A., Smith, R. J., & Macmillan, D. C. (2014). Evaluating conservation flagships and flagship fleets. Conservation Letters, 7 (3), 263–270. doi: 10.1111/conl.12070. Veríssimo, D., Fraser, I., Groombridge, J., Bristol, R., & Macmillan, D. C. (2009). Birds as tourism flagship species: a case study of tropical islands Research methods. Animal Conservation, 12 (6), 549–558. doi: 10.1111/j.1469-1795.2009.00282.x. Veríssimo, D., MacMillan, D. C., & Smith, R. J. (2011). Toward a systematic approach for identifying conservation flagships. Conservation Letters, 4, 1–8. doi: 10.1111/j.1755-263X.2010.00151.x. Veríssimo, D., Pongiluppi, T., Santos, M. C. M., Develey, P. F., Fraser, I., …, & Macmilan, D. C. (2013). Using a systematic approach to select flagship species for bird conservation. Conservation Biology, 28, 269–277. doi: 10.1111/cobi.12142. Veríssimo, D., Vaughan, G., Ridout, M., Waterman, C., Macmillan, D., & Smith, R. J. (2017). Increased conservation marketing effort has major fundraising benefits for even the least popular species. Biological Conservation, 211, 95–101. doi: 10.1016/j.biocon.2017.04.018. UNEP-WCMC., & IUCN. (2018). Protected Planet: The World Database on Protected Areas (WDPA). Cambridge, UK: UNEP-WCMC and IUCN. Available at: http://www.protectedplanet.net (accessed 21 August 2018). United Nations Educational, Scientific, and Cultural Organization (UNESCO). (2003). Towards a strategy for the sustainable development of tourism in the Sahara, in the context of combating poverty. The Sahara of Cultures and People. Paris, France. Waldron, A., Mooers, A. O., Miller, D. C., Nibbelink, N., Redding, D., & Kuhn, T. S. (2013). Targeting global conservation funding to limit immediate biodiversity declines. Proceedings of the National Academy of Sciences, 110 (29), 1–5. doi: 10.1073/pnas.1221370110. Walpole, M., L., & Leader-Williams, N. (2002). Tourism and flagship species in conservation. Biodiversity and Conservation, 11 (3), 43-547. doi: 10.1023/A:1014864708777. Wright, A. J., Veríssimo, D., Pilfold, K., Parsons, E. C. M., Ventre, K., Cousins, J., …, McKinley, E. (2015). Competitive outreach in the 21st century: Why we need

104 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

conservation marketing. Ocean & Coastal Management, 115, 41-48. doi: 10.1016/j.ocecoaman.2015.06.029. Weaver, D. (2001). The Encyclopedia of Ecotourism. Wallingford, UK: CABI Publishing.

FCUP 105 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Chapter 4

Mapping ecotourism and other cultural ecosystem services

Article III Santarém, F., Campos, J.C., Pereira, P., Hamidou, D., Saarinen, J., Brito, J.C. (2018). “Using multivariate statistics to assess ecotourism potential of waterbodies: A case-study in Mauritania”. Tourism Management, 67: 34-46

Article IV Santarém, F., Saarinen, J., Brito, J.C. (2020). “Mapping and analysing cultural ecosystem services in conflict areas”. Ecological Indicators, 110: 105943.

106 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Article III: Using multivariate statistics to assess ecotourism potential of waterbodies: A case-study in Mauritania.

Frederico Santarém1,2, João Carlos Campos1,2, Paulo Pereira1,2, Dieng Hamidou3, Jarkko Saarinen4,5, José Carlos Brito1,2

1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 – Faculté des Sciences et Techniques, Université des Sciences, de Technologie et de Médecine de Nouakchott, Nouakchott, Mauritania. 4 – Geography Research Unit, University of Oulu, Finland 5 – School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

Abstract

Evaluating the ecotourism potential of sites is a key issue in tourism management. Multiple methodologies have been developed to assess the ecotourism potential of sites. However, there are many constraints affecting their quality. Methodologies independent of subjective criteria and weights are lacking, compromising following interpretations on where to allocate efforts for ecotourism development. We propose a new approach to circumvent these issues that combines independent statistical procedures to assess ecotourism potential. By combining multi-criteria with ordination and clustering algorithms, this two-stage statistical approach allowed identifying suitable water-bodies for ecotourism development in Mauritania and independently assessed which features are related with ecotourism potential. The method was able to group sites for different ecotourist demands, which has implications for policy makers and tourism planners trying to optimize investments while protecting biodiversity and supporting communities. We provide a framework that is scalable and applicable by stakeholders operating in ecotourism planning and management worldwide.

FCUP 107 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Keywords tourism planning; deserts; multi-criteria approach; ordination methods; clustering algorithms; low-income countries; wetlands.

108 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Introduction

Ecotourism is considered an efficient tool for biodiversity conservation (Fennell, 2015; Lindsey, Alexander, du Toit, & Mills, 2005; Mossaz, Buckley, & Castley, 2015; Penteriani et al., 2017). It helps building environmental awareness among ecotourists and local communities (Honey, 2008; Reimer & Walter, 2013) and can fund effectively conservation initiatives (Kirkby et al., 2010, 2011), resulting in several positive net-effects of environmental outcomes (see Bricker & Kerstetter, 2017; Buckley, 2009; and Weaver, 2002). Ecotourism may be a potential tool for the reduction of poverty in low-income countries (Goal 1 of the United Nations Sustainable Development Goals) and the sustainable economic development of poor communities with decent employment for all (Goal 8; http://www.un.org/sustainabledevelopment/sustainable-development-goals/; see Reimer & Walter, 2013; Saarinen & Rogerson, 2013; Scheyvens, 1999). Despite potential negative impacts, such as marginalization of local communities and species habituation to human presence (see Penteriani et al., 2017; Weaver, 2002), ecotourism is widely being promoted as a strategy that fosters beneficial relationships between tourism, the environment and local communities living in low-income countries, especially in Africa (Lindsey et al., 2005; United Nations, 2017). The potential of ecotourism in contributing to poverty alleviation and threatened species preservation has been claimed for African deserts (Santarém & Paiva, 2015; Weaver, 2001), particularly for the Sahara-Sahel ecoregions (Brito et al., 2014, 2016). All countries in the region are low-income nations, most of them exhibiting low human development index (UNDP, 2016). Historically, they have been underfunded for poverty alleviation and biodiversity loss retention schemes (Durant et al., 2012; Waldron et al., 2013). Mauritania is one of these African arid countries, where around 40% of people live in rural areas, lacking both infrastructural and health conditions (FAO, 2013). The country displays multiple cultural, biological and landscape values that may be suitable for sustaining ecotourism projects (Vale, Pimm, & Brito, 2015). Mauritanian water-bodies are of particularly relevance, as they are rich in human heritage and biodiversity. On the one hand, they may be surrounded by well-preserved villages where traditional livelihoods and secular ceremonies still occur. There are ancient wells for cattle watering, evidences of intense intellectual activity propagated by caravans crossing the ksours (Mauritanian towns) since the Middle Ages, and archaeological remains in the form of rock-paintings (Hosni, 2000; UNESCO, 2003). On the other hand, water-bodies may hold up to 78% of the endemic fishes, amphibians, reptiles and mammals of Mauritania (Vale et al., 2015). They gather large concentrations of water-birds(Cooper, Shine, McCann, FCUP 109 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

& Tidane, 2006), that provide bird-watching opportunities. The surrounding habitats, where dunes intersect with unique desert-adapted agricultural patches (oasis), and the presence of geological formations (waterfalls) on rock canyons are of great interest for ecotourism. In addition, several water-bodies contain relict West African crocodile populations (Brito, Martínez-Freiría, Sierra, Sillero, & Tarroso, 2011) that have been considered as a national flagship species (Telleria, Ghaillani, Fernandez-Palacios, Bartolome, & Montiano, 2008), which supports the development of species-based conservation and ecotourism programmes (Veríssimo et al., 2013, 2014). Based on these values, the potential for ecotourism development in the desert water-bodies of Mauritania is promising, but requires urgent assessment and research- informed planning, given the local societal development and biodiversity conservation needs. One potential caveat to developing such assessments in these regions is related to the lack of a systematic approach to evaluate the ecotourism potential of sites with knowledge-based and well-informed criteria for ecotourism planning. This paper proposes a new method that aims to go beyond subjective criteria and weights in ecotourism site planning by combining independent statistical procedures to assess ecotourism potential. By combining multi-criteria with ordination and clustering algorithms, the two-stage statistical approach is used here to identify suitable water- bodies for ecotourism development and to assess independently which features are related with ecotourism site-potential.

Approaches to evaluate ecotourism potential

Ecotourism is a relatively new concept in tourism planning but its basic elements and dimensions have been discussed for decades in studies on tourism and natural resource management (Fennell, 2015; Honey, 2008; Stronza, 2007; Weaver, 2001). There are numerous definitions for ecotourism (see Donohoe & Needham, 2006). In general, it refers to responsible travel to natural areas with an aim to use tourism demand for conservation goals, environmental education and local communities benefits (Diamantis, 1999;TIES, 2017). There is a strong focus on sustainable tourism development with an emphasis on ecologically sound and driven practices (Saarinen, 2006). Therefore, comprehensive biodiversity evaluations and assessments are often included to ecotourism planning. In this context, multi-criteria approaches combining ecological and social data, and spatial analyses conducted in Geographical Information System (GIS) environments have been integrated to assess the ecotourism potential of many sites around the globe.

110 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

The multi-spatial layers and the continuous analyses allowed by GIS tools increase the spatial accuracy of predictions and the information handled, thus assisting in the decision-making process (Dhami, Deng, Burns, & Pierskalla, 2014; Dhami, Deng, Strager, & Conley, 2017). For instance, GIS and participatory methods (Delphi and Analytic Hierarchy Process, AHP) were used to map ecotourism as a cultural service (Nahuelhual et al., 2013), high resolution IKONOS images and GIS were used for planning useful resources to ecotourism (Fung & Wong, 2007), and visitor’s preferences were assessed for mapping weighted and unweighted potential sites for forest-based ecotourism development (Dhami et al., 2014). Despite the plethora of methods being proposed for assessing cultural ecosystem services (where ecotourism is included), there are still constraints affecting the quality of those assessments. Most studies rely on a priori controlled approaches that require weighting of variables, either via inquiries or expert-knowledge scoring. While they are often practical, such approaches have also been criticized (Kliskey, 2000), and methods that are independent of subjective criteria and weights are still lacking (Milcu, Hanspach, Abson, & Fischer, 2013). Circumventing a priori classification approaches requires independent methods that assess the contribution of variables without necessarily weighting them, such as ordination methods and clustering algorithms. Principal Component Analysis (PCA) is a statistical procedure that uses orthogonal transformations to convert observations of possible correlated variables into linearly uncorrelated variables, which emphasizes variation and brings out relevant patterns in a dataset (Hotelling, 1933). PCA produces a series of orthogonal uncorrelated axes, where the first component accounts the largest possible variability in the data and the following components account for decreasing proportions of the variability. Cluster analysis is an unsupervised learning task in data mining that consists in grouping a set of objects in a way that objects in the same group (called a cluster) are more similar to each other than to those in other groups (Tryon, 1939). Both statistical approaches have been applied independently in tourism research. For instance, PCA was used to assess social preferences for recreation demands in a three-stage procedure for mapping recreation potential (Peña et al., 2015) and to identify suitable variables associated to the implementation of recreation activities (Kliskey, 2000), while cluster analysis was used to identify distinct groups of adventure travellers (Sung, 2004). The advantage of using PCA is that it allows the empirical quantification of the influence of biological, ecological, and human-related variables in the suitability of a site for ecotourism development. Cluster analysis allows to group sites sharing similar characteristics, which can help managing distinct groups of ecotourists with particular FCUP 111 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel demands. When used jointly with ordination methods, the two-stage multivariate statistics approach improves interpretations in tourism studies (Arimond & Elfessi, 2001; Pina & Delfa, 2005). The application of multivariate statistics to assess suitable sites for recreation development has been emphasised by some authors (e.g. Kliskey, 2000), but the use of ordination methods and clustering algorithms for pairwise grouping remains unexplored in ecotourism research (Weaver, 2001). This combination of statistical methods might help elucidating where to allocate efforts to meet different ecotourists demands, especially in regions offering conditions to attend both ‘hard’ (environmentally specialized, physically challenging, minimally commodities serviced, strongly biocentric) and ‘soft’ (shallow interaction with nature, physically passive, comfort demanding, heavily commodities serviced, shallow interaction with nature) ecotourist profile needs (Fennell & Weaver, 2005; Weaver, 2001, 2002). Here, we combine for the first time a PCA with a clustering algorithm to identify suitable locations for ecotourism development. We explore approaches based in multi- criteria coupled with ordination and clustering algorithms that are independent from a priori weighting. Specifically, we want to address the following questions: 1) Which variables contribute the most to explain the variability of water-bodies to ecotourism development?; 2) How many groups of water-bodies can be distinguished according to their characteristics for ecotourism potential?; and 3) Which groups of water-bodies display characteristics associated with potential for ecotourism development? We aim to demonstrate that the methodological process is suitable to characterize the ecotourism potential of regions with recreation appeal that lack conservation and development funding, and that it can be scalable and replicable worldwide.

Methods

Study area

Mauritania is located between 15.8°N and 20.6°N and west of 9.5°W. Climate is characterized by a dry, cool season approximately from November to February and a dry, hot season approximately from March to June. Annual rains are scarce and seasonal, with most precipitation occurring between August and September (Cooper et al., 2006). The country has a poor road system and key infrastructures are almost only available in main cities (WEF, 2017). The methodological process was applied to a set of 30 water-bodies dispersed throughout the southern mountains of Tagant, Assaba and Afollé (Fig. 3.1). These water-

112 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

bodies display distinct characteristics in topography, altitude, landscapes, and other environmental and anthropogenic features, comprising gueltas (rock-pools in mountainous canyons), tâmoûrts (seasonal floodplains located on the foothills of the mountains), sources, lakes, and oueds (rivers; see details about water-bodies definitions in Brito et al., 2011). The water-bodies were visited by four researchers during 22 days in January 2016. The coordinates of water-bodies were gathered from a Global Positioning System (GPS) and were displayed in ArcGIS 10.1 (WGS84 datum).

Fig. 3.1. Location of the 30 water-bodies studied (and code number), main settlements with supporting infrastructures (and their names), paved roads, rivers, and mountains. The location of the capital (Nouakchott) is depicted in the small inset.

The multi-criteria approach

A total of 16 variables were selected for analyses, including two variables related to the biological interest of water-bodies, nine variables describing their environmental traits, and four variables depicting human activities and infra-structures associated to them (Table 3.1). FCUP 113 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table 3.1. Types of variables used to assess the ecotourism potential of water-bodies, their codes, description, interest for ecotourism (Min-minimum and Max-maximum), units, and source. Type Code Description Interest Units Source Biological CR Maximum number of Max Number Brito et al., 2011; Campos, Martínez-Freiría, Sousa, crocodiles reported Santarém, & Brito, 2016 BI Bird species richness Max Number Fieldwork index Environmental AREA Surface area Max m2 Fieldwork SL Topographic Max Degrees SRTM 90m Digital Elevation Database v4.1 heterogeneity (http://www.cgiar-csi.org) WA Water availability Max Adimensional Campos, Sillero, & Brito 2012 NDVI Vegetation Max Adimensional LP DAAC (http://lpdaac.usgs.gov) Productivity NWB Euclidian distance to Min Meters Fieldwork and GIS analysis the nearest water- body WBH Habitat heterogeneity Max Number PostelMedias: GLOBCOVER 2006 (http://postel.obs- mip.fr) RH Habitat heterogeneity Max Number PostelMedias: GLOBCOVER 2006 (http://postel.obs- from the nearest mip.fr) paved road

114 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

DA Area covered by Max m2 PostelMedias: GLOBCOVER 2006 (http://postel.obs- dunes mip.fr) AA Area covered by Max m2 PostelMedias: GLOBCOVER 2006 (http://postel.obs- agricultural land mip.fr) MTWM Maximum temperature Min ◦C Worldclim (http://www.worldclim.org) of the warmest month Anthropogenic DR Distance to the Min Meters Fieldwork and GIS analysis nearest paved road DI Distance to the Min Meters Fieldwork and GIS analysis nearest settlement with supporting infrastructures IUCN Number of IUCN Min Number Campos, Martínez-Freiría, Sousa, Santarém, & Brito, threat factors 2016 categories HII Human influence Min Adimensional SEDAC (http://sedac.ciesin.columbia.edu) index FCUP 115 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

The 16 variables were selected based on established relationships with the ecotourism potential of the water-bodies to attract international tourists (Table 3.1). Flagship species are considered key for ecotourism development (Veríssimo, Fraser, Groombridge, Bristol, & MacMillan, 2009; Walpole & Leader-Williams, 2002). Thus, we selected widely accepted flagship species, such as crocodiles (Telleria et al., 2008) and birds (Veríssimo et al., 2009). The 30 water-bodies were sampled to record the maximum number of individual West African crocodiles (Crocodylus suchus) and the number of different bird species. The sampling effort was on average 0.273 man/hours in each water-body. Estimations of crocodile numbers followed standard sampling protocols (Brito et al., 2011; Campos, Martínez-Freiría, Sousa, Santarém, & Brito, 2016). In water- bodies where crocodiles were not detected in January 2016, but were known to occur, the number of estimates were based in previous reports (Brito et al., 2011). The number of bird species in each water-body was accessed by point sampling, using binoculars and telescopes. Species identification was based in field-guides (Borrow & Demey, 2001). The number of observed species was weighted by the sampling effort in each water-body (in minutes) to obtain an index of bird species richness that considers temporal biases in sampling efforts and allows sites to be compared. Environmental features, such as high slopes may influence ecotourism potential of a site (Dhami et al., 2014, 2017), and we measured it by calculating the standard deviation of the slope (derived from altitude at 90m resolution) in 1x1km window around each water-body, using the “Buffers” tool of ArcGIS 10.1 (ESRI, 2012). Water-bodies with large areas and permanent water availability play a crucial role in desert environments, allowing the observation of large species diversity and traditional human activities, such as water pumping from mudbrick-wells for watering cattle (Brito et al., 2014, 2016; Hosni, 2000; Vale et al., 2015). Thus, the area of each water-body was estimated in the field with a GPS and the water availability was assessed using the Normalized Difference Water Index (NDWI) at 1 arc second resolution in a 1x1km window around water-bodies. Vegetation heterogeneity is well appreciated by tourists (Chhetri & Arrowsmith, 2008; Dhami et al., 2014, 2017) and thus, we used the Normalized Difference Vegetation Index (NDVI) time-series from the period between 2000 and 2016, at 250m resolution to extract it in a 1x1km window around water-bodies. Short geographic distances between visiting sites are best acknowledged because tourists and ecotourism enterprises tend to maximize, in general, the return on time/money invested to visit a region (Kienast, Degenhardt, Weilenmann, Wäger, & Buchecker, 2012). Thus, we calculated the Euclidean distances between water-bodies. High habitat heterogeneity influences a site appraisal for ecotourism (Nahuelhual et al., 2013; Santarém, Silva, & Santos, 2015). Thus, we recorded the habitat heterogeneity

116 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

around water-bodies and from the nearest paved road to the water-bodies, using a 1x1km window around water-bodies. In desert environments, the observation of dunes adds an increased experience to the observer (Cooper et al., 2006; Hosni, 2000). We calculated the area covered by dunes in a 1x1km window surrounding the water-bodies based on land-cover classification by remote sensing. Local agriculture is mostly based in ancient practices that may add value to the aesthetic visit of water-bodies (Chan, Shaw, Cameron, Underwood, & Daily, 2006). Thus, we calculated the area covered by agricultural patches in a 1x1km window surrounding water-bodies. Weather conditions can influence recreation opportunities and in remote regions they are of great concern for the well-being of tourists (Kienast et al., 2012). Thus, we extracted the maximum temperature of the warmest month in a 1x1km window surrounding the water-bodies. Anthropogenic factors, such as the time to access a touristic point or a settlement with supporting infrastructures, may be crucial for tourism activities (Castro, Souza, & Thapa, 2015; Kienast et al., 2012). Time to get to a visiting site is greatly influenced by the distance and the road type (Chan et al., 2006; Paracchini et al., 2014; Wong & Fung, 2015). Thus, we measured the distance from the water-bodies to the nearest settlement with supporting infrastructures (hospitals, restaurants or accommodation resources; Chhetri & Arrowsmith, 2008). As most of the water-bodies are far from the main paved roads (see Fig. 3.1), we also measured the distance from the water-bodies to the nearest paved road (Dhami et al., 2014). Both distances were recorded using a GPS. Anthropogenic impacts on environment are global drivers of ecological processes that influence biodiversity as a whole (Sanderson et al., 2002) and affect biodiversity availability for tourism experiences in the field. The number of threats affecting the water- bodies was quantified following the IUCN Threats Classification Scheme Version 3.2 (www.iucnredlist.org/technicaldocuments/classification-schemes/threats- classificationscheme) and was available from Campos et al. (2016). The Human Influence Index (HII) was extracted in a 1x1km window around each water-body.

Data analysis

We used a PCA to identify which variables contribute the most to explain the variability of water-bodies for ecotourism development. To do so, we scaled data by subtracting each variable value by its mean and then dividing by its standard deviation. This process converted each raw datum score into a standardized value with mean 0 and standard deviation 1. We then plotted water-bodies along the two first principal components to identify main patterns of clustering of water-bodies. FCUP 117 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Groups of water-bodies sharing similar biological, environmental and anthropogenic characteristics for ecotourism were identified using a model-based clustering approach. Ten multivariate normal mixture models with different parameterizations concerning the distribution, volume, shape and orientation of the covariance structure of the multivariate data of the water-bodies were estimated by maximum likelihood using an expectation-maximization algorithm. The best model and the most likely number of clusters were chosen using the Bayesian Information Criterion (BIC). Analyses were performed using the three first, most explaining, principal components to assess the number of optimum clusters. Finally, we contrasted the distribution of the groups identified by the clustering algorithm with the variables represented in the first and second principal component (PC1 and PC2, respectively) axes. This process allowed identifying groups of water-bodies suitable for different sets of ecotourists, ‘hard’ and ‘soft’ (Fennell & Weaver, 2005; Weaver, 2001, 2002). Both the PCA and the clustering analysis were performed in R environment (R Core Team, 2016) with the mclust package (Fraley & Raftery, 2002; Fraley, Raftery, Murphy, & Scrucca, 2012).

Results

Contribution of variables to the variability of water-bodies for ecotourism development

The variables with the highest influence in the first, second and third axes of the PCA were: water availability (WA) and maximum temperature of the warmest month (MTWM) in the first axis; vegetation productivity (NDVI) and distance to the nearest settlement with supporting infrastructures (DI) in the second axis; and distance to the nearest paved road (DR) and habitat heterogeneity from the nearest paved road to water-bodies (RH) in the third axis (Table 3.2). The two first axes accounted for 38% of the variability. The representation of water-bodies along the two first axes allowed identifying three groups of water-bodies for ecotourism development (Fig. 3.2).

Grouping of water-bodies according to ecotourism potential

We used the model with the highest BIC (spherical, equal volume with seven components, “EII”; Fig. B1) to group water-bodies according to their most similar characteristics for ecotourism. The BIC increased with the number of clusters and most

118 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

of the increase occurred with less than seven clusters, which were chosen as the optimum number of clusters defining the groups for ecotourism development. These seven subgroups of water-bodies fall inside the three main groups of water-bodies identified by the PCA (Fig. 3.2).

Suitability of water-bodies for ecotourism development

The green group defined by the PCA contained three subgroups of water-bodies defined by the clustering algorithm (Fig. 3.2) that, in comparison to other groups, tended to display more flagship species (larger crocodile populations, more bird species), higher habitat heterogeneity around the water-body, higher water availability, larger flat surface areas, shorter distances between each other, wider agriculture areas in the surroundings, and the highest human impacts (Table B1; Figs. 3.3 and B2). The blue PCA group gathered water-bodies that tended to exhibit the highest topographic heterogeneity and habitat heterogeneity along the roads to reach water-bodies, the highest temperatures in the warmest month, the largest distances to paved roads and cities with supporting infra-structures, the lowest productivity, and the fewer number of IUCN threats (Table B1; Figures 3.3 and B2). The red PCA group contained three BIC subgroups of water-bodies that tended to display the highest topographic heterogeneity, the fewer number of flagship species, and the smallest areas and water availability. Within this group, there is a subgroup (dark red squares) of water-bodies that recorded numerous IUCN threat factors and are far from paved roads and infrastructures (Table B1; Fig. 3.3 and B2). Overall, there is a trend for lack of geographic structure in the distribution of the groups, but the main blue group (Fig. 3.2) is restricted to the northernmost water-bodies of the Tagant mountain and subgroup A2 is restricted to the eastern Afollé mountain (Fig. 3.3). FCUP 119 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

1 Table 3.2. Loadings of the principal components, and the standard deviation (SD) and cumulative variance (CUM) of the principal components. Variable PC1 PC2 PC3 PC4 PC5 PC6 PC7 PC8 PC9 PC10 PC11 PC12 PC13 PC14 PC15 PC16 CR -0.166 -0.061 0.039 0.631 -0.227 -0.043 0.129 -0.043 -0.279 -0.052 0.502 -0.231 -0.246 0.144 -0.160 0.059 BI -0.314 -0.184 -0.013 -0.115 -0.250 -0.116 0.398 -0.604 0.158 -0.250 -0.059 0.310 -0.020 -0.198 -0.122 0.118 AREA -0.291 -0.073 0.193 -0.295 -0.184 -0.377 0.177 0.562 -0.139 -0.025 0.068 0.108 0.144 -0.010 -0.418 -0.176 SL 0.254 0.215 0.124 0.011 0.508 -0.283 0.331 -0.207 0.184 -0.076 -0.088 -0.245 -0.130 0.282 -0.406 -0.120 WA -0.458 0.093 -0.060 -0.210 0.142 0.059 0.146 -0.157 -0.016 0.177 0.150 -0.634 0.273 -0.225 0.201 -0.208 NDVI 0.050 -0.493 0.200 -0.137 -0.088 -0.029 0.033 0.098 0.004 0.133 -0.346 -0.416 -0.526 -0.168 -0.004 0.248 NWB -0.312 -0.057 0.224 -0.249 0.193 0.350 -0.433 -0.061 0.101 -0.429 0.206 0.026 -0.328 0.143 -0.154 -0.218 WBH -0.190 -0.211 -0.048 0.111 0.500 0.304 0.383 0.232 -0.369 -0.283 -0.166 0.131 0.129 0.058 0.132 0.250 RH -0.136 0.204 -0.484 0.121 -0.115 -0.079 -0.193 -0.072 -0.388 -0.204 -0.524 -0.110 -0.150 -0.078 -0.212 -0.277 AA -0.354 -0.137 0.201 0.295 0.095 0.233 -0.120 -0.119 0.100 0.587 -0.308 0.183 0.134 0.192 -0.294 -0.104 DA -0.176 -0.082 -0.432 0.035 0.422 -0.259 -0.044 0.110 0.144 0.283 0.264 0.284 -0.396 -0.323 0.032 -0.045 MTWM -0.411 0.272 -0.139 -0.083 -0.108 -0.228 0.006 0.111 0.211 0.009 -0.138 -0.028 -0.182 0.581 0.316 0.346 DR 0.043 -0.320 -0.498 -0.060 -0.003 0.083 -0.203 0.036 0.254 -0.090 0.125 -0.223 0.334 0.123 -0.434 0.377 DI 0.134 -0.456 -0.291 -0.054 -0.140 0.100 0.273 0.039 0.118 0.020 0.029 0.017 -0.051 0.413 0.219 -0.587 IUCN 0.033 -0.329 0.092 -0.178 0.187 -0.456 -0.337 -0.348 -0.496 0.088 0.092 0.053 0.161 0.248 0.136 0.053 HII -0.121 -0.228 0.169 0.465 0.113 -0.372 -0.212 0.121 0.383 -0.364 -0.193 -0.082 0.236 -0.159 0.225 -0.153 SD 1.771 1.670 1.468 1.250 1.143 1.028 0.876 0.833 0.741 0.648 0.590 0.495 0.423 0.395 0.267 0.166 CUM 0.203 0.383 0.523 0.624 0.708 0.776 0.826 0.871 0.906 0.934 0.956 0.972 0.984 0.994 0.998 1.000 2

120 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 3.2. Water-bodies grouped according to the three first components of the Principal Component Analysis (PCA) and Bayesian Information Criterion (BIC). The seven subgroups (blue triangles; dark green, green and light green circles; dark red, red and orange squares) identified by the BIC clustering are within the three groups (blue, green and red hulls) identified by the PCA. FCUP 121 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 3.3. Distribution of the three groups identified by the Principal Component Analysis (PCA; squares, circles and triangles) and of the seven subgroups identified by the Bayesian Information Criterion (BIC) models (different colours of squares, circles and triangles). See Table B1 for grouping details and Figure B2 for grouping examples.

Discussion

By combining ordination and clustering algorithms, we could identify sites for ecotourism development in a systematic way, by emphasizing strong patterns across different groups of water-bodies. The method allowed evaluating independently which variables contributed the most for the definition of ecotourism potential of each site. To our knowledge, this is the first study doing that. The methodology here presented has important implications for decision-making in ecotourism planning and can help stakeholders identifying the best sites for local investment. This two-step approach can be scalable and replicable to other regions where ecotourism interest is growing, and where proper consultation and management are needed to avoid further impacts. We propose a set of improvements to be considered if this methodology is to be applied in the future.

122 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Methodological improvements from previous approaches

The methodology proposed herein improved the identification of suitable sites for ecotourism. The main advantage of this method in detriment to previous approaches is the combination of ordination (PCA) and clustering algorithms (BIC) in a non-heuristic way. The ordination method allowed assessing the contribution of variables without requiring expert knowledge, which otherwise could compromise following interpretations due to subjectivity of people’s choices (Milcu et al., 2013). The cluster analysis allowed grouping sites sharing similar characteristics for ecotourism, which can help managing distinct groups of ecotourists with particular demands. Also, by estimating the parameters of the mixture models with empirical data, the choice of optimum number of clusters was a statistical decision, rather than an intuitive judgment. This approach avoided choosing the optimum number of clusters heuristically, as in non-model-based approaches (Fraley & Raftery, 2002). This approach assessed the contribution of variables without necessarily weighting them before further analyses, thus circumventing a priori classification schemes. To our knowledge, this was the first study combining ordination methods and clustering algorithms for pairwise grouping in ecotourism research, which addresses previous criticisms on dependent methods relying on subjective criteria and expert-knowledge scoring (Kliskey, 2000). This process avoids misleading interpretations, thus maximizing the investment planning towards the places suitable for ecotourism development. The multi-criteria approach here used constitutes further strength of the methodology. It relied on key biological, environmental and anthropogenic features to assess the ecotourism potential of water-bodies. This approach considers both the ecotourism attraction capacity, i.e. the natural features of the territory, and the demand side, i.e. the conditions that influence the tourism attraction capacity of the sites. The combination of both sets of attributes is crucial for ecotourism assessments and remained under-explored to date (Nahuelhual et al., 2013).

Methodological aspects needing further research

Despite this two-stage multivariate statistics approach allowed circumventing previous caveats in tourism studies, some methodological improvements shall be considered to refine its application. In this study, only continuous variables were considered for the PCA input. However, assessments of ecotourism potential often compile categorical variables, which may be beneficial to characterise the suitability of sites. Statistics such as Principal Coordinates Analysis (PCoA), FCUP 123 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel allow exploring data dissimilarities and can analyse jointly categorical and continuous variables. This multidimensional scaling method starts with a dissimilarity matrix (distance matrix) and assigns for each entry in the matrix a location in a low-dimensional space. PCoA tries to find the main axes through a matrix and calculates a series of eigenvalues and eigenvectors, in a similar way as PCA does (Pavoine, Dufour, & Chessel, 2004). For instance, PCoA has been explored for pairwise grouping of Sahara-Sahel endemic species (Vale & Brito, 2015). Its potential for tourism studies remains unexplored. The multi-criteria approach here used considered an array of distinct variables related to ecotourism potential of the water-bodies. They were suitable for the scale of the sites under analysis (isolated points in space), but ecotourism potential assessments can consider other variables that are dependent on continuous scales. For instance, other animal groups could also be included in the analyses, after evaluating which flagship species may appeal to broader audiences (Veríssimo et al., 2009; Veríssimo, MacMillan, & Smith, 2011). When performing analysis at broader scales, other biological groups may also be considered (Santarém et al., 2015). For instance, southern Mauritania concentrates the largest proportion of species richness of the country (Brito et al., 2016), and thus water-bodies located in these regions might be of special interest for ecotourists. The economic welfare of water-bodies nearest villages and the local willingness of villagers for ecotourism investments should be also considered in assessments (Castro et al., 2015), as investment priority should be given to those communities with fewer resources. Furthermore, increasing the representativeness of sampling may result in better performance of the model-based clustering algorithm, which may elucidate where to allocate efforts for ecotourism implementation.

Factors relevant for assessing suitability of desert water-bodies for ecotourism

The multi-statistics approach here used allowed the identification of key variables to assess ecotourism potential of desert water-bodies, such as water availability. Duration of standing water was previously found to be key for wildlife in desert contexts (Cooper et al., 2006), and greater areas with permanent water usually concentrate more species that appeal to the ecotourist, especially endemic species with high flagship potential (Vale et al., 2015). Particularly, they aggregate flocks of birds, highly appreciated by ecotourists and birdwatchers worldwide (Veríssimo et al., 2009; Veríssimo et al., 2014) that are frequently willing to spend considerable amounts of money to spot these flagship species (Steven, Castley, & Buckley, 2013). Thus, water availability should be considered in future assessments of ecotourism potential of water-bodies.

124 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Maximum temperature of the warmest month and vegetation were also identified as being associated with suitability of water-bodies to ecotourism. Despite high temperatures in the peak of the hot season were not found to be relevant for recreation in temperate environments (Switzerland; Kienast et al., 2012), they may be considered in ecotourism activities in desert environments (e.g. in Mauritania). Likewise, vegetation is known to be an influencing factor for ecotourism development in non-desert biomes (Dhami et al., 2014, 2017) and a key predictor of scenic attractiveness in an Australian region with savannahs and shrub lands (Chhetri & Arrowsmith, 2008), supporting its consideration in recreation and ecotourism potential assessments in all biomes. The potential observation of multiple habitats along the roads leading to the recreational sites was identified as influencing the ecotourism potential of water-bodies. This observation is supported by previous studies that found the number of different habitats in a line-transect influencing landscape appeal (Santarém et al., 2015). Particularly, agricultural patches around desert water-bodies add an increased value to the ecotourist. As ancient practices of agriculture in arid environments are reflected around these water-bodies (Cooper et al., 2006), agricultural land should be considered when assessing the potential of desert water-bodies for ecotourism development. The methodology explored herein found both distance-related variables (to roads and infra-structures) to be associated to the ecotourism potential of a site, which is particularly relevant for desert contexts (Weaver, 2001). Remoteness was recently found to be one of the most important criteria for ecotourism development for both visitors and ecotourism-experts (Dhami et al., 2017). Though Paracchini et al. (2014) found a positive correlation between remoteness level and recreation opportunity in Europe, in desert contexts, such as the Sahara-Sahel, characterised by poor road systems and hostile weather/environments (WEF, 2017), planners may wish to consider distance as a pivotal constrain to ecotourism development. Intermediate levels of distances to roads and supporting infrastructures can be considered, depending on ecotourists demands and territorial contexts (Nahuelhual et al., 2013).

Traits of suitable water-bodies for ecotourism development

The PCA discriminated three core groups of water-bodies with distinct traits that may be suitable for ecotourists with different demands. This has implications for ecotourism planners that want to maximize return-on-investment and that need to consider the demands of distinct hard- and soft-types of ecotourists (Weaver, 2001, 2005). One group of water-bodies (green hull; Figs. FCUP 125 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

3.2-3.3) exhibits several characteristics that may be ideal for the soft-ecotourist, which usually demands short trips with physical comfort and nearby services, and has moderate environmental commitment (Weaver, 2001, 2005). These water-bodies are in proximity to cities with supporting infrastructures and display large water-surface areas with high water availability, where flagship species concentrate. The northernmost group of water-bodies (blue hull; Figs. 3.2-3.3) recorded the highest air temperatures and topographic variations, and thus may be challenging to visit by the most common ecotourist. These water-bodies are apparently suitable for “hard-desert” ecotourists trying to obtain the highest return from the desert natural landscapes visited. Hard-ecotourists are usually looking for physical challenging environments, travel in small and specialized groups to regions with few if any services available, and have strong environmental commitment (Weaver, 2001, 2005). The third group of water-bodies (red hull; Figs. 3.2-3.3) displays characteristics that are apparently less suitable for ecotourism development. These include low water-availability, low agricultural and dunes-covered lands, and low number of flagship species, alongside the long distances from roads and supporting infrastructures. These water-bodies exhibit numerous IUCN threats, which may have implications for both biodiversity and ecotourists pursuing natural experiences in deserts. From the ten water-bodies identified as biodiversity hotspots by Vale et al. (2015), we have here accessed half of them, and four out of these five water-bodies were identified to be less suitable for ecotourism development. The water-bodies within this group should be prioritized for reducing current threat levels before considering them for ecotourism development.

Conclusions

Here, a two-stage multivariate statistic was developed, which allowed grouping water-bodies according to their different characteristics for ecotourism development. Such grouping has implications for ecotourism planners and academics that need to tailor tourist visits according to the profiles of different ecotourists. As ecotourism planners may face different demands from distinct ecotourist types, this methodology helps maximizing return-on-investments and enhancing local economies. However, as these biodiversity rich sites often involve intrinsic cultural heritage values and local livelihoods, ethical considerations involved with ecotourism, developments and related research needs to be carefully considered. The proposed approach does not indicate where we should but where we could place tourism activities in an optimal

126 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

situation. Thus, any approach to develop ecotourism in these water-bodies needs to be further contextualized and underline their specificities, as emphasised in this paper, to not deteriorate these fragile socio-ecological systems in touristic uses. By adding ecological and anthropogenic variables to biological data of water-bodies, the methodology presented here can provide a basic information where to optimally allocate investments for wildlife conservation and local development, while meeting different ecotourist demands. This is especially relevant in low- income countries and poor peripheral regions with high tourism potential that lack economic resources and conservation and ecotourism investments. In this respect, policy makers might use the framework here provided to achieve the United Nations Sustainable Goals, particularly in what respect combating poverty while protecting biodiversity and key ecosystem services. FCUP 127 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

References Arimond, G., & Elfessi, A. (2001). A clustering method for categorical data in tourism market segmentation research. Journal of Travel Research, 39(4), 391–397. https://doi.org/10.1177/004728750103900405 Borrow, N., & Demey, R. (2001). Birds of Western Africa. London: Christopher Helm. Bricker, K. S. & Kerstetter, D. L. (2017). Effecting positive change – an introduction. Journal of Ecotourism, 16(3), 201-202. https://doi.org/10.1080/14724049.2017.1390936 Brito, J. C., Godinho, R., Martínez-Freiría, F., Pleguezuelos, J. M., Rebelo, H., Santos, X., … Carranza, S. (2014). Unravelling biodiversity, evolution and threats to conservation in the Sahara-Sahel. Biological Reviews, 89(1), 215–231. https://doi.org/10.1111/brv.12049 Brito, J. C., Martínez-Freiría, F., Sierra, P., Sillero, N., & Tarroso, P. (2011). Crocodiles in the Sahara desert: An update of distribution, habitats and population status for conservation planning in Mauritania. PLoS ONE, 6(2), e14734. https://doi.org/10.1371/journal.pone.0014734 Brito, J. C., Tarroso, P., Vale, C. G., Martínez-Freiría, F., Boratyński, Z., Campos, J., … Carvalho, S. B. (2016). Conservation Biogeography of the Sahara-Sahel: Additional protected areas are needed to secure unique biodiversity. Diversity and Distributions, 22, 371–384. https://doi.org/10.1111/ddi.12416 Buckley, R. (2009). Evaluating the net effects of ecotourism on the environment : A framework, first assessment and future research. Journal of Sustainable Tourism, 17(6), 643–672. https://doi.org/10.1080/09669580902999188 Campos, J. C., Martínez-Freiría, F., Sousa, F. V., Santarém, F., & Brito, J. C. (2016). Update of distribution , habitats , population size , and threat factors for the West African crocodile in Mauritania. Amphibia-Reptilia, 37(3), 325–330. https://doi.org/10.1163/15685381-00003059 Campos, J. C., Sillero, N., Brito, J. C. (2012). Normalized difference water indexes have dissimilar performances in detecting seasonal and permanent water in the Sahara–Sahel transition zone. Journal of Hydrology, 464-465, 438-446. https://doi.org/10.1016/j.jhydrol.2012.07.042 Castro, E. V. De, Souza, T. B., & Thapa, B. (2015). Determinants of Tourism Attractiveness in the National Parks of Brazil. Parks, 21(2), 51–62. Chan, K. M. A., Shaw, M. R., Cameron, D. R., Underwood, E. C., & Daily, G. C. (2006). Conservation planning for ecosystem services. PLoS Biology, 4(11), 2138–2152. https://doi.org/10.1371/journal.pbio.0040379 Chhetri, P., & Arrowsmith, C. (2008). GIS-based Modelling of Recreational Potential of Nature- Based Tourist Destinations. Tourism Geographies, 10(2), 233–257.

128 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

https://doi.org/10.1080/14616680802000089 Cooper, A., Shine, T., McCann, T., & Tidane, D. A. (2006). An ecological basis for sustainable land use of Eastern Mauritanian wetlands. Journal of Arid Environments, 67, 116–141. https://doi.org/10.1016/j.jaridenv.2006.02.003 Dhami, I., Deng, J., Burns, R. C., & Pierskalla, C. (2014). Identifying and mapping forest-based ecotourism areas in West Virginia - Incorporating visitors’ preferences. Tourism Management, 42, 165–176. https://doi.org/10.1016/j.tourman.2013.11.007 Dhami, I., Deng, J., Strager, M., & Conley, J. (2017). Suitability-sensitivity analysis of nature- based tourism using geographic information systems and analytic hierarchy process. Journal of Ecotourism, 16(1), 41–68. https://doi.org/10.1080/14724049.2016.1193186 Diamantis, D. (1999). The Concept of Ecotourism : Evolution and Trends. Current Issues in Tourism, 2(2–3), 93–122. https://doi.org/10.1080/13683509908667847 Donohoe, H. M., & Needham, R. D. (2006). Ecotourism: The Evolving Contemporary Definition. Journal of Ecotourism, 5(3), 192–210. https://doi.org/10.2167/joe152.0 Durant, S. M., Pettorelli, N., Bashir, S., Woodroffe, R., Wacher, T., Ornellas, P. de, … Baillie, J. (2012). Forgotten Biodiversity in Desert Ecosystems. Science, 336(6087), 1379–1380. https://doi.org/10.1126/science.336.6087.1379 Environmental Systems Research Institute (ESRI). (2012). ArcGIS Release 10.1. Redlands, CAFood and Agriculture Organization of the United Nations (FAO). (2013). Low-Income Food-Deficit Countries (LIFDC) — List for 2013. Retrieved February 23, 2017, from http://www.fao.org/countryprofiles/lifdc/en/FAO 2011 Fennell, D. A. (2001). A Content Analysis of Ecotourism Definitions. Current Issues in Tourism, 4(5), 403–421. https://doi.org/10.1080/13683500108667896 Fennell, D. A. (2015). Ecotourism (4th ed.). Oxon and New York: Routledge. Fennell, D., & Weaver, D. (2005). The Ecotourium Concept and Tourism-Conservation Symbiosis. Journal of Sustainable Tourism, 13(4), 373–390. https://doi.org/10.1080/09669580508668563 Fraley, C., & Raftery, A. (2002). Model-based clustering , discriminant analysis , and density estimation. Journal of the American Statistical Association, 97(458), 611–631. https://doi.org/10.1198/016214502760047131 Fraley, C., Raftery, A., Murphy, T. B., & Scrucca, L. (2012). mclust Version 4 for R : Normal Mixture Modeling for Model-Based Clustering , Classification , and Density Estimation. Fung, T., & Wong, F. K. K. (2007). Ecotourism planning using multiple criteria evaluation with GIS. Geocarto International, 22(2), 87–105. https://doi.org/10.1080/10106040701207332 FCUP 129 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Honey, M. (2008). Ecotourism and sustainable development: Who owns paradise? (2nd Ed.). Washington DC: Island Press. Hosni, E. (2000). Strategy for sustainable tourism development in the Sahara. Paris: UNESCO. Hotelling, H. (1933). Analysis of a complex of statistical variables into principal components. Journal of Educational Psychology, 24(6), 417–441. https://doi.org/10.1037/h0071325 Kienast, F., Degenhardt, B., Weilenmann, B., Wäger, Y., & Buchecker, M. (2012). GIS-assisted mapping of landscape suitability for nearby recreation. Landscape and Urban Planning, 105(4), 385–399. https://doi.org/10.1016/j.landurbplan.2012.01.015 Kirkby, C. A., Giudice-Granados, R., Day, B., Turner, K., Velarde-Andrade, L. M., Dueñas- Dueñas, A., … Yu, D. W. (2010). The market triumph of ecotourism: An economic investigation of the private and social benefits of competing land uses in the Peruvian Amazon. PLoS ONE, 5(9), 22–27. https://doi.org/10.1371/journal.pone.0013015 Kirkby, C. A., Giudice, R., Day, B., Turner, K., Soares-Filho, B. S., Oliveira-Rodrigues, H., & Yu, D. W. (2011). Closing the ecotourism-conservation loop in the Peruvian Amazon. Environmental Conservation, 38(1), 6–17. https://doi.org/10.1017/S0376892911000099 Kliskey, A. D. (2000). Recreation terrain suitability mapping: A spatially explicit methodology for determining recreation potential for resource use assessment. Landscape and Urban Planning, 52(1), 33–43. https://doi.org/10.1016/S0169-2046(00)00111-0 Lindsey, P. A., Alexander, R. R., du Toit, J. T., & Mills, M. G. L. (2005). The potential contribution of ecotourism to African wild dog Lycaon pictus conservation in South Africa. Biological Conservation, 123(3), 339–348. https://doi.org/10.1016/j.biocon.2004.12.002 Milcu, A. I., Hanspach, J., Abson, D., & Fischer, J. (2013). Cultural ecosystem services: A literature review and prospects for future research. Ecology & Society, 18(3), 44–88. https://doi.org/10.5751/ES-05790-180344 Mossaz, A. C., Buckley, R., & Castley, J. G. (2015). Ecotourism Contributions to Conservation of African Big Cats. Journal for Nature Conservation, 28, 112–118. https://doi.org/10.1016/j.jnc.2015.09.009 Nahuelhual, L., Carmona, A., Lozada, P., Jaramillo, A., & Aguayo, M. (2013). Mapping recreation and ecotourism as a cultural ecosystem service: An application at the local level in Southern Chile. Applied Geography, 40, 71–82. https://doi.org/10.1016/j.apgeog.2012.12.004 Paracchini, M. L., Zulian, G., Kopperoinen, L., Maes, J., Schägner, J. P., Termansen, M., … Bidoglio, G. (2014). Mapping cultural ecosystem services: A framework to assess the potential for outdoor recreation across the EU. Ecological Indicators, 45, 371–385. https://doi.org/10.1016/j.ecolind.2014.04.018

130 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Pavoine, S., Dufour, A. B., & Chessel, D. (2004). From dissimilarities among species to dissimilarities among communities: A double principal coordinate analysis. Journal of Theoretical Biology, 228(4), 523–537. https://doi.org/10.1016/j.jtbi.2004.02.014 Peña, L., Casado-Arzuaga, I., & Onaindia, M. (2015). Mapping recreation supply and demand using an ecological and a social evaluation approach. Ecosystem Services, 13, 108–118. https://doi.org/10.1016/j.ecoser.2014.12.008 Penteriani, V., López-Bao, J. V., Bettega, C., Dalerum, F., Delgado, M. del M., Jerina, K., … Ordiz, A. (2017). Consequences of brown bear viewing tourism: A review. Biological Conservation, 206, 169–180. https://doi.org/10.1016/j.biocon.2016.12.035 Pina, I. P. A., & Delfa, M. T. D. (2005). Rural tourism demand by type of accommodation. Tourism Management, 26(6), 951–959. https://doi.org/10.1016/j.tourman.2004.06.013 R Core Team. (2016). R: A Language and Environment for Statistical Computing. Retrieved from https://www.r-project.org/ Reimer, J. K., & Walter, P. (2013). How do you know it when you see it? Community-based ecotourism in the Cardamom Mountains of southwestern Cambodia. Tourism Management, 34, 122–132. https://doi.org/10.1016/j.tourman.2012.04.002 Saarinen, J. (2006). Traditions of Sustainability in Tourism Studies. Annals of Tourism Research, 33(4), 1121–1140. https://doi.org/10.1016/j.annals.2006.06.007 Saarinen, J., & Rogerson, C. M. (2013). Tourism and the Millennium Development Goals: perspectives beyond 2015. Tourism Geographies, 16(1), 23–30. https://doi.org/10.1080/14616688.2013.851269 Sanderson, E. W., Jaiteh, M., Levy, M. A., Redford, K. H., Wannebo, A. V, & Woolmer, G. (2002). The Human Footprint and the Last of the Wild. BioScience, 52(10), 891–904. https://doi.org/https://doi.org/10.1641/0006-3568(2002)052[0891:THFATL]2.0.CO;2 Santarém, F., & Paiva, F. (2015). Conserving desert biodiversity through ecotourism. Tourism Management Perspectives, 16, 176–178. https://doi.org/10.1016/j.tmp.2015.07.016 Santarém, F., Silva, R., & Santos, P. (2015). Assessing ecotourism potential of hiking trails: A framework to incorporate ecological and cultural features and seasonality. Tourism Management Perspectives, 16, 190–206. https://doi.org/10.1016/j.tmp.2015.07.019 Scheyvens, R. (1999). Ecotourism and the empowerment of local communities. Tourism Management, 20(2), 245–249. https://doi.org/10.1016/S0261-5177(98)00069-7 Steven, R., Castley, J. G., & Buckley, R. (2013). Tourism Revenue as a Conservation Tool for Threatened Birds in Protected Areas. PLoS ONE, 8(5), 1–8. https://doi.org/10.1371/journal.pone.0062598 FCUP 131 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Stronza, A. (2007). The Economic Promise of Ecotourism for Conservation. Journal of Ecotourism, 6(3), 210–230. https://doi.org/10.2167/joe177.0 Sung, H. H. (2004). Classification of Adventure Travelers: Behavior, Decision Making, and Target Markets. Journal of Travel Research, 42(4), 343–356. https://doi.org/10.1177/0047287504263028 Telleria, J. L., Ghaillani, H. E. M., Fernandez-Palacios, J. M., Bartolome, J., & Montiano, E. (2008). Crocodiles Crocodylus niloticus as a focal species for conserving water resources in Mauritanian Sahara. Oryx, 42(2), 292–295. https://doi.org/10.1017/s0030605308007850 The International Ecotourism Society (TIES). (2015). What is ecotourism? Retrieved December 04, 2017, from www.ecotourism.org/what-is-ecotourism. Tryon, R. C. (1939). Cluster analysis: Correlation profile and orthometric (factor) analysis for the isolation of unities in mind and personality. Michigan: Edwards brother, Incorporated, lithoprinters and publishers. United Nations. (2017). Economic Development in Africa report 2017: Tourism for transformative and inclusive growth. Retrieved December 04, 2017, from: http://unctad.org/en/PublicationsLibrary/aldcafrica2017_en.pdf United Nations Development Programme (UNDP). (2016). Human Development Report 2016. Human Development for Everyone. New York, USA. United Nations Educational, Scientific, and Cultural Organization (UNESCO). (2003). Towards a strategy for the sustainable development of tourism in the Sahara, in the context of combating poverty. The Sahara of Cultures and People. Paris, France. Vale, C. G., & Brito, J. C. (2015). Desert-adapted species are vulnerable to climate change: Insights from the warmest region on Earth. Global Ecology and Conservation, 4, 369–379. https://doi.org/10.1016/j.gecco.2015.07.012 Vale, C. G., Pimm, S. L., & Brito, J. C. (2015). Overlooked mountain rock pools in deserts are critical local hotspots of biodiversity. PLoS ONE, 10(2), e0118367. https://doi.org/10.1371/journal.pone.0118367 Veríssimo, D., Fraser, I., Gir, W., Campos, A. A., Smith, R. J., & Macmillan, D. C. (2014). Evaluating Conservation Flagships and Flagship Fleets. Conservation Letters, 7(3), 263– 270. https://doi.org/10.1111/conl.12070 Veríssimo, D., Fraser, I., Groombridge, J., Bristol, R., & Macmillan, D. C. (2009). Birds as tourism flagship species: a case study of tropical islands Research methods. Animal Conservation, 12, 549–558. https://doi.org/10.1111/j.1469-1795.2009.00282.x Veríssimo, D., MacMillan, D. C., & Smith, R. J. (2011). Toward a systematic approach for

132 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

identifying conservation flagships. Conservation Letters, 4(1), 1–8. https://doi.org/10.1111/j.1755-263X.2010.00151.x Veríssimo, D., Pongiluppi, T., Santos, M. C. M., Develey, P. F., Fraser, I., Smith, R. J., & Macmilan, D. C. (2013). Using a Systematic Approach to Select Flagship Species for Bird Conservation. Conservation Biology, 28(1), 269–277. https://doi.org/10.1111/cobi.12142 Waldron, A., Mooers, A. O., Miller, D. C., Nibbelink, N., Redding, D., & Kuhn, T. S. (2013). Targeting global conservation funding to limit immediate biodiversity declines. Proceedings of the National Academy of Sciences, 110(29), 1–5. https://doi.org/10.5061/dryad.p69t1 Walpole, M. J., & Leader-Williams, N. (2002). Tourism and flagship species in conservation. Biodiversity and Conservation, 11, 543–547. https://doi.org/10.1023/A:1014864708777 Weaver, D. B. (2001). The encyclopedia of ecotourism. Wallingford: CABI. Weaver, D. B. (2002). The evolving concept of ecotourism and its potential impacts. International Journal of Sustainable Development, 5(3), 251–264. https://doi.org/10.1504/IJSD.2002.003753 Weaver, D. B. (2005). Comprehensive and minimalist dimensions of ecotourism. Annals of Tourism Research, 32(2), 439–455. https://doi.org/10.1016/j.annals.2004.08.003 World Economic Forum (WEF). (2017). The Travel & Tourism Competitiveness Report 2017. Retrieved September 18, from: https://www.weforum.org/reports/the-travel-tourism- competitiveness-report-2017. Wong, F. K. K., & Fung, T. (2015). Ecotourism planning in Lantau Island using multiple criteria decision analysis with geographic information system. Environment and Planning B: Planning and Design, 43(4), 640–662. https://doi.org/10.1177/0265813515618583

FCUP 133 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Article IV: Mapping and analysing cultural ecosystem services in conflict areas

Frederico Santarém1,2,3, Jarkko Saarinen3,4, José Carlos Brito1,2

1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 - Geography Research Unit, University of Oulu, Finland 4 - School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

Abstract

Human-mediated global environmental change threatens ecosystem services worldwide. Detailed cultural ecosystem services mapping is crucial to counteract ecosystem degradation, but such mapping exercises have been confined to small-scale analyses in developed countries. Additionally, disturbances constraining the supply of cultural ecosystem services transboundary have never been mapped, which hampers the accurate management of ecosystems, particularly in underdeveloped countries affected by human conflicts. The Sahara-Sahel ecoregions of Africa represent an excellent model to map the distribution of transboundary attractions and constraints to cultural ecosystem services due to the many conflicts affecting its drylands. We mapped and analysed the supply of cultural ecosystem services in the Sahara-Sahel, using a multicriteria approach that includes transboundary attractions and constraints playing at broader scales. We wanted to understand where are located the hotspots of cultural ecosystem services and which regions displaying the highest levels of attractions may be simultaneously threatened by constraint features. Overall, 35.4% of the study area displays high (27.9%) to very high (7.5%) levels of attractions to cultural ecosystem services supply, while 8.6% of the area displays high (7.5%) to very high (1.1%) levels of constraints that limit the usufruct of these services in the region. Our findings showed that the main mountains and wetlands are supplying high levels of cultural ecosystem services but are threatened in some parts of their range by transboundary constraints. Some country-borders displayed a high concentration of constraints impacting desert biodiversity and human communities. This highlights the urgency of policymakers to reinforce

134 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

transboundary strategic actions to halt the ongoing destruction of natural resources in the region. The developed approach is scalable and replicable in any ecosystem, including in those located in data-scarce regions. Including constraints to ecosystem services supply is paramount to achieve the United Nations Sustainable Development Goals.

Keywords Mapping ecosystem services; cultural services; deserts; drylands; Sahara; Sahel; ecotourism; constraints; ecotourism attractions; Sustainable Development Goals FCUP 135 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Introduction

Global environmental change and unsustainable human activities affect ecosystems worldwide and decision-makers struggle to create policies that halt biodiversity loss and related human impoverishment (Díaz et al., 2019). The idea of ecosystem services has been increasingly adopted to manage and counteract the degradation of ecosystems worldwide, to safeguard the natural capital for future generations, to provide new opportunities for local economies, and to highlight the direct and indirect contributions of ecosystems to human wellbeing (Costanza et al., 2014; Martinez-Harms et al., 2015; Kubiszewski et al., 2017). Recently, within the overall ecosystem service thinking, cultural ecosystem services (CES) have received an increasing importance in research and policy making (TEEB, 2010; Ståhhammar and Pedersen, 2017). The Millennium Ecosystem Assessment (MEA, 2005) defines cultural ecosystem services as the non- material benefits people obtain from ecosystems through spiritual enrichment, cognitive development, reflection, recreation, and aesthetic experience, and which include aesthetic, spiritual, educational, and recreational services (Daniel et al., 2012; Milcu et al., 2013; Stålhammar and Pedersen, 2017). CES have been recognised as crucial to human wellbeing (Safriel and Adeel, 2005) and, despite recent critics on their operationalisation and value to land- management (see Kirchhoff, 2019), are now part of the strategies to achieve many of the targets under the United Nations Sustainable Development Goals (UN-SDGs), such as SDG1 – no poverty, SDG8 – economic growth, and SDG15 – life on land (UNWTO and UNDP, 2017; Wood et al. 2018). Particularly, ecotourism and related recreational activities have been proposed as a key tool to frame programmes aiming to eradicate extreme poverty and to stimulate environmental protection in fragile areas (UNWTO, 2018a). To achieve the UN-SDGs, it is crucial that decision-makers contemplate the actual spatial characteristics of CES supply (i.e., the capacity of an area to provide ecosystem goods within a given time; Burkhard et al., 2012a; Lu et al., 2018) in land-use planning decisions (Maes et al., 2012; Schulp et al., 2014; Peña et al., 2015). Reviews on ecosystem services mapping (e.g., Martínez-Harms and Balvanera, 2012; Crossman et al., 2013) have revealed that CES are among the least commonly mapped ecosystem services, which highlights the need for additional research. Indeed, mapping, and analysing CES are important to identify priority areas ensuring long-term provision of ecosystem services (Martínez-Harms and Balvanera, 2012; Lu et al., 2018), but they may be challenging due to their intangible nature and dependence in social models (Daniel et al., 2012; Milcu et al., 2013; Small et al., 2017). Some studies have attempted to map and analysed CES supply worldwide, with a bias towards developed countries (Milcu et

136 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

al., 2013). However, most of the studies are limited to local (from 100 to 1000 km2) and regional (from 1000 to 100,000 km2) geographical scales (e.g. Kienast et al., 2012; Nahuelhual et al., 2013; Plieninger et al., 2013; Casado-Arzuaga et al., 2014; van Berkel and Verburg, 2014; Peña et al., 2015; Gigović et al., 2016; and Nahuelhual et al., 2017), while very few have tried to cover continental or global scales (>1,000,000 km2; Martinez-Harms et al., 2015; but see Paracchini et al., 2014; van Zanten et al., 2016; and Komossa et al., 2018). Mapping CES supply at continental or sub-continental scales, however, supports decision- making and implementation in multiple ways. First, it helps identifying priorities for ecosystem management and restoration across political borders, which may enhance the preservation of natural assets through the development of transboundary "green" infrastructure networks that benefit neighbouring societies (Naidoo et al., 2008; Schulp et al., 2014). Second, it allows addressing thoroughly the continuum of ecosystem services across multiple scales, as many CES tend to be omni-directional, flowing to beneficiaries in many directions (Rosa et al., 2017). Third, it helps considering and analysing environmental problems that do not fit into the context of small spatial scales, avoiding overlooking ecological and social processes that operate across large scales of management (Martinez-Harms et al., 2015; Rosa et al., 2017). Forth, it allows cross- cultural comparisons of various recreation values that are being transformed by similar processes across the globe. Lastly, it helps understanding the diversity and similarities of continent’s cultural- historical backgrounds on desiring the preservation of iconic landscapes (van Zanten et al., 2016). Mapping CES supply requires extensive information to capture the heterogeneity of recreational attractions and constraints in ecotourism (Egoh et al., 2007; Komossa et al., 2018). As such, CES mapping using multicriteria approaches linked with Geographical Information Systems (GIS) has increased exponentially over the last years (Wolff et al., 2015). However, most analyses use proxy variables (e.g., land-cover) instead of primary CES data (Seppelt et al., 2011; Martínez-Harms and Balvanera, 2012; Crossman et al., 2013), despite the potential biases induced by proxies (Eigenbrod et al., 2010; Cerretelli et al., 2018). Mapping CES supply is usually based in coupling recreation values into integrative maps after weighting criteria via participatory methods (e.g., Kienast et al., 2012; Nahuelhual et al., 2013; Plieninger et al., 2013; van Berkel and Verburg, 2014; Peña et al., 2015; Nahuelhual et al., 2017; Komossa et al., 2018). In this respect, participatory methods are useful to understand how potential ecotourists valorise different landscape attributes (van Berkel and Verburg, 2014), but it can also include several methodological limitations: 1) they can only be performed in moderately to highly visited places; 2) they are subject to biases, for instance on choosing the stakeholders to engage in the study and on stakeholders overvaluing the most well-known places; 3) they cannot assure the complete FCUP 137 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel commitment of interest groups for participation; 4) they are expensive to conduct; and 5) they present spatial and temporal limitations that hamper the systematic comparison of results, especially across large political relevant scales (van Berkel and Verburg 2014; Brown and Fagerholm, 2015; Wolff et al., 2015; Small et al., 2017; Komossa, et al., 2018; Scholte et al., 2018; Wood et al., 2018). In addition, mapping and analysing land disturbances threatening CES supply is considered crucial to conserve the biodiversity underpinning those same services (O’Farrell et al., 2010), yet, most CES studies fall short to include representative constraint factors causing changes in CES delivery at broader geographical scales (Martinez-Harms et al., 2015; Hanaček and Rodríguez-Labajos, 2018). For instance, land-degradation is decreasing the capacity of ecosystems to provide ecosystem services at unprecedented rates, urging the identification of priority areas for intervention (Cerretelli et al., 2018). Constraints to CES supply, such as regional insecurity caused by conflict acting across country borders (e.g., attacks by extremist groups in several adjacent African countries; Harmon, 2016; Brito et al., 2018) or by mining expansion (Hanaček and Rodríguez-Labajos, 2018), emphasise the need to address CES mapping at transboundary scales. Mapping transboundary constraints to CES is of paramount importance for policymakers to identify areas where changes are impacting ecosystems, classify avoidance-areas for ecotourists (Lanouar and Goaied, 2019), and allocate recreation resources thoroughly (Lu et al., 2018). In remote and poorly visited regions, alternative approaches independent from participatory methods are needed to assess the attractions and the constraints to CES supply at continental scales. The Sahara-Sahel ecoregions of Africa represent an excellent model to test alternative methodologies to map and analyse transboundary CES attractions and constraints. On the one hand, the Sahara-Sahel is mostly dominated by deserts and arid landscapes (Dinerstein et al., 2017), which constitute one of the last wild biomes on Earth (Watson et al., 2018) and offers prospects for ecotourists seeking ‘last-chance-to-see’ wild places (see Lemelin et al., 2012; Saarinen, 2018). In the current context of global environmental change, this kind of last chance tourism (LCT) refers to tourism motivated by a need to experience a place or certain environmental condition before they may disappear (Hall and Saarinen, 2010; Piggott-McKellar and McNamara, 2016). In this respect, Sahara-Sahel is still rich (although under threat) in unique biodiversity adapted to aridity that can be found nowhere else in the world (Brito et al., 2014; Durant et al., 2014; Brito and Pleguezuelos, 2019) and provide important ecosystem services for local people. Examples of these services include water-supply (crucial for settling people near water-rich regions) and food- and medicines-supply of desert-adapted plants (e.g. Nitraria retusa and Herniaria hirsuta, respectively), as well as regulating services like pollination by local wildlife,

138 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

and supporting services such like sand fixation by particular desert-adapted plants (Safriel & Adeel, 2005; Davies et al., 2012; Bidak et al., 2015; Ghazi et al., 2018). In particular, mountains are rich in endemic and flagship species (Brito et al., 2016; Santarém et al., 2019a) inhabiting small and relict wetlands of global importance (Vale et al., 2015) and in geological, cultural and historical features that help understanding how human communities adapted to arid environments throughout time (Santarém et al., 2019b). These desert attractions have been partially explored in CES mapping at local scales (e.g., in Mauritanian wetlands; Santarém et al., 2018), but standardised CES mapping at transboundary level is lacking. On the other hand, the Sahara-Sahel spreads across 18 countries, most of them ranked among the least human developed (UNDP, 2018) and least visited in the world (UNWTO, 2018b), which hampers the development of participatory methods to weight recreation values. For instance, Mali and Chad are among the poorest countries in the world and among the 30 least visited countries by international tourists. Some of these countries also account for the largest reported tourist deaths associated to conflicts and terrorism (Dioko and Harrill, 2019). Regional insecurity, such as transboundary attacks, kidnapping and smuggling routes, has contributed to the impoverishment of local people and the depletion of natural resources that international ecotourists would enjoy if they were preserved in first place (Lanouar and Goaied, 2019). For instance, Sudan, Libya and Mali have recorded an increased number of conflict events since 2011 and some borders (e.g., Mali-Algeria, Mali-Niger) are now under control by armed groups or terrorists (Brito et al., 2018). Constraints to CES supply have been partially mapped in the Sahara- Sahel (Brito et al., 2018), but others, such as pipelines of natural resources exploitation or landmines, remain unmapped. In addition, some countries (e.g. Morocco and Libya) may face the strongest species’ habitat-suitability losses that will force a large proportion of species to be up- listed in conservation status (Powers and Jetz, 2019), but remain mostly neglected for biological conservation (Durant et al., 2012; Waldron et al., 2013; Brito et al., 2014; Durant et al., 2014; Brito & Pleguezuelos, 2019) and CES mapping exercises (Seppelt et al., 2011; Hanaček and Rodríguez-Labajos, 2018). Consequently, our understanding of which Sahara-Sahel regions concentrate the largest environmental and cultural attractions and which transboundary features constraint CES supply at larger scales is very deficient. Mapping and analysing attractions and constraints to the supply of CES in the Sahara-Sahel is needed to help decision-makers developing strategies that improve the benefits of these services to human well-being, while avoiding conflict-areas (Hanaček & Rodríguez-Labajos, 2018). The main objective of this work is to identify, map and analyse the supply of CES, specifically for recreation and ecotourism, in the Sahara-Sahel. The study utilises a multicriteria FCUP 139 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel approach that considers primary and secondary data, including constraints playing at broader scales, and that is independent from participatory methods (see Fig. 3.4 for details on the workflow). In particular, we want to understand: 1) which is the distribution of individual attraction and constraint features?; 2) which regions within the Sahara-Sahel concentrate the largest number of attractions and constraints for CES supply?; 3) given the spatial variability in attractions and constraints, which regions concentrate the highest levels of attractions and simultaneously display or not constraint features? Given the inherent characteristics of the Sahara-Sahel, we hypothesise that most of the attractions for CES supply will be located in the main mountains and wetlands of the region, and that most of the constraints will tend to cluster in the same regions where constraints have been estimated to locate in previous studies (see Brito et al., 2014, 2018).

140 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 3.4. Flowchart detailing the steps taken to map and analyse cultural ecosystem services supply in the Sahara-Sahel. FCUP 141 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Methods

Study area The study area (Fig. 3.5a; Fig. C1) comprises the Sahara, the largest warm desert in the world, and the contiguous arid Sahel (≈11,200,000 km2). To map and analyse CES supply, we used as baseline the ecoregion limits (Dinerstein et al., 2017) and divided it in 4120 pixels of 0.5- degree resolution (WGS84 coordinate reference system).

Distribution of attraction and constraint features A multicriteria approach was applied to map and analyse CES supply, based on variables that are known to condition the decision of potential ecotourists to visit regions (see below). A total of 26 variables were considered, including primary and secondary data organised in two categories: 21 attraction and five constraint features (see Tables 3.3, C1 and C2 for data sources and details on variables).

Table 3.3. Categories and subcategories of the variables assessed to map and analyse cultural ecosystem services in the Sahara- Sahel, their code, original data types and data sources. * See Table C1 for details. Categories Code Variable Original Data sources data type Attractions Biodiversity Speci Total species Polygon Brito et al. (2016); IUCN (2018)

Endem Endemic richness Polygon Brito et al. (2016); IUCN (2018)

Flags Flagship richness Polygon IUCN (2018); Santarém et al. (2019)

Ecore Terrestrial ecoregions Polygon Dinerstein et al. (2017) Conservation Fores Forest reserves Point NGA (2016)

ProtA Protected Areas Polygon IUCN and UNEP-WCMC (2017)

Herit UNESCO World Point UNESCO (2018) Heritage Sites Landscape LandF Major landscape Polygon GeoCover (2018), updated features* from PSSC (2018)

142 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

GorMt Gorges and mountain Point Updated from NGA (2016) passes

RockF Peculiar rock Point Updated from NGA (2016) formations

Caves Caves Point NGA (2016)

Wetla Major wetlands* Polygon GeoCover (2018)

Guelt Rock pools (known as Point Updated from Brito et al. guelta) (2014) Cultural EthnG Major ethnographic Polygon Updated from OECD-SWAC groups* (2014)

Oases Oases Point Updated from NGA (2016)

Monum Monuments Point Updated from NGA (2016) Historical CarVi Caravan villages Point Updated from OECD-SWAC (2014)

CarRo Caravan routes* Polyline Updated from OECD-SWAC (2014)

Forti Fortifications from Point Updated from NGA (2016) colonial period

Ruins Ruins, tombs, sites of Point Updated from NGA (2016) empires historical land occupation

RockA Rock art Point Updated from Lluch and Philip (2003); Gauthier and Gauthier (2006); Jesse et al. (2007); Gauthier and Gauthier (2008); Le Quellec (2009); Riemer (2009); Gauthier and Gauthier (2011); Noguera and Zboray (2011); Biagetti et al. (2013); Gallinaro (2013); Le Quellec (2013); Barnett and Gaugnin (2014); Brémont (2018) Constraints Conflict Landm Landmines Point DesertInfo (2006) FCUP 143 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Attac Attacks/battles and Point de Hass (2007); Ewi (2010), violence against Raleigh et al. (2010); civilians (2011-2016) Beauchamp (2014a,b); Grossman (2015); START (2015); Weiss (2016); Brito et al. (2018)

Migra Smuggling and human Polyline Brachet et al. (2011); migration routes Rekacewicz (2012); OECD- SWAC (2014); Brito et al. (2018) Exploitation ExtraF Oil, gas, mining Point NIMA (1997); Duncan et al. of natural extractive facilities (2014) resources Pipel Pipelines Polyline NIMA (1997)

Attraction features Animal species are one of the main attraction elements to a given site and localities exhibiting high species richness are linked with increased recreational value (Chung et al., 2018). Importantly, endemic and flagship species are key factors to attract ecotourists (Santarém et al., 2019a) and have been used in mapping CES supply (e.g., Scholte et al., 2018). Thus, we considered the total number of amphibian, , bird and mammal species occurring in the Sahara-Sahel (Fig. 3.5b), and the number of endemic (Fig. 3.5c) and flagship species (Fig. 3.5d) of these groups. We quantified species richness of these three groups per pixel based on polygons depicting individual species ranges. The species richness of endemics by pixel was weighted over the total species found in each pixel. Ecoregions, i.e., ecological regionalisations that delineate areas of similar environmental conditions, ecological processes and biotic communities (Dinerstein et al., 2017), help distinguishing different landscape determinants of CES supply (Weyland and Laterra, 2014). Forest Reserves and other protected areas are crucial for biodiversity preservation and nature recreation worldwide (Balmford et al., 2015; Chung et al., 2018) and they have been assessed in other CES studies (Scholte et al., 2018). UNESCO World Heritage Sites help preserve the cultural and natural heritage of universal importance and provide an additional value to ecotourists seeking places of outstanding value (Levin et al., 2019). Based on this we quantified the number of ecoregions (Fig. 3.5e), forest reserves, protected areas and UNESCO sites (Fig. 3.5f) per pixel.

144 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 3.5. Map of the Sahara-Sahel (a) and distribution of the biodiversity (b-e) and conservation features (f) in the region. BF – Burkina Faso; CAM – Cameroon; CAR – Central African Republic; ER – Eritrea; ETH – Ethiopia; NGA – Nigeria; SEN – Senegal; SSD – South Sudan.

In deserts, specific landscape features provide multiple opportunities for recreation. For instance, sand dunes and flat plains provide scenic settings that trigger deep emotional connections to nature and provide solitude and tranquillity from stressful urbanised cores (Cooper et al., 2006; Santarém et al., 2018, 2019b). Continental and coastal cliffs, as well as gorges and mountain passes create wonder to landscape aficionados. Rock canyons, valleys, inselbergs, escarpments, and rock formations (such as rock arches, desert potholes, aeolian landforms, and FCUP 145 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel saltpans) and caves are also prominent desert landscape attractions to recreationists and ecotourists. High altitude areas and associated plateaus, volcanic cones and meteor impact craters are also geological features of potential interest to desert ecotourists (UNEP, 2006a,b; Santarém et al., 2019b), thus, we calculated the number of these landscape features per pixel (Figs. 3.6a-e). Wetlands are crucial for species and humans in deserts and tend to concentrate large species diversity and traditional human activities (Brito et al., 2014; UNEP, 2016a) and have been assessed in other CES studies (Plieninger et al., 2013; Peña et al., 2015; Scholte et al., 2018). Particularly, mountain rock-pools constitute local biodiversity hotspots (Vale et al., 2015) and are fundamental to ecotourism activities in deserts (UNEP, 2016b; Santarém et al., 2018). We quantified the number of wetlands and mountain rock-pools per pixel (Fig. 3.6f).

146 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 3.6. Distribution of major landscape features (a-c), caves and gueltas (d), gorges and mountain passes (e) and wetlands and gueltas (f) in the Sahara-Sahel.

Historical and cultural features associated with human presence in deserts add value to the general recreational experience. Indigenous people have long adapted to the harsh arid environment by using desert elements to counter extreme temperatures and winds and a generalised paucity of water (UNEP, 2006b; Santarém et al., 2019b). The presence and diversity of ethnic groups has been linked with ecotourism potential (Zeppel, 2006; Fennell, 2015; Saarinen et al., 2014; Saarinen 2016). We counted the number of ethnic groups per pixel (Fig. 3.7a). Oases represent an example of cultural adaptation to arid environments. They were created and managed by local communities for agriculture purposes (mostly for production of date FCUP 147 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel palms, Phoenix dactylifera) that benefit from the unpredictability of rain. They are now a vital asset to CES supply that benefit local people living in their surroundings (Davies et al., 2012), mostly because of their ecological and cultural importance (UNEP, 2006a,b; Santarém et al., 2018, 2019b). Grounded on this, we quantified the number of oases per pixel (Fig. 3.7b). The historical occupation of deserts is noticeable in cultural monuments, which comprise large structures erected to commemorate notable persons or events and are valued among many recreationists (van Berkel and Verburg 2014; Peña et al., 2015). We counted the number of monuments per pixel (Fig. 3.7b). Commercial villages and caravan routes have long been established in deserts for trading exchange and religious-cultural enrichment. They offer extensive depositories of desert knowledge and are highly appreciated by ecotourists (UNEP, 2006a; Santarém et al., 2019b). Thus, we quantified the number of caravan villages and routes per pixel (Fig. 3.7c). Furthermore, many historical sites have been identified for desert CES (UNEP, 2006a,b; Santarém et al., 2019b). In Sahara-Sahel, fortifications from colonial periods and ruins, tombs and sites of historical empire land occupation (e.g., Roman and Ancient Egypt) abound (UNEP, 2006a). Therefore, we quantified the number of fortifications and ruins per pixel (Fig. 3.7d). Finally, rock paintings and engravings illustrate past livelihoods and past climatic shifts in the regions where they are depicted. In particular, the Sahara Desert is considered an open book of Human history (Gallinaro, 2013). Here, desert rock art is highly recognizable among international recreationists and ecotourists because it is easily found in large concentrations (Santarém et al., 2019b). We quantified the number of rock art sites per pixel (Fig. 3.7e).

148 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 3.7. Distribution of cultural (a-b) and historical features (c-e) in the Sahara-Sahel.

Constraint features Accelerated land-use changes and unsustainable development programmes are striking deserts. Preliminary constraining elements to desert recreation have been identified in the Sahara-Sahel, such as armed conflicts, military-defensive structures (e.g., landmines) and violence against civilians (Brito et al., 2018; Santarém et al., 2019b). Smuggling and human migration are now a common issue in many parts of the region (OECD-SWAC, 2014; Harmon, 2016). Natural resources exploitation, via oil, gas and mining facilities, is increasing in the Sahara- Sahel, threatening local biodiversity (Brito et al., 2014; Duncan et al., 2014; Brito et al., 2018) and CES supply (Hanaček and Rodríguez-Labajos, 2018). These constraints deplete the desert from FCUP 149 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel natural and cultural features (Brito et al., 2014; Levin et al., 2019), and impact the recreation potential of the region. We quantified the number of these constraining features per pixel (Figs. 3.8a-d).

Fig. 3.8. Distribution of conflict (a-c) and exploitation of natural resources features (d) in the Sahara-Sahel.

Identifying hotspots of attractions and constraints to cultural ecosystem services supply To map and analyse the locations of CES hotspots in the Sahara-Sahel we first standardised each of the 26 variables, dividing each entry value in the database by its maximum value, to obtain re-scaled variables varying between 0 and 1 (as in Peña et al., 2015). Then, the 26 standardised variables were mapped in ArcMap 10.1 (ESRI, 2012) to obtain the individual distribution of standardised attractions and constraints to CES supply in the study area. After, regions remarkably supplying higher levels of CES were identified by combining the 21 standardised attraction features to produce a map of CES attractions hotspots, and the same procedure was employed to map the location of constraints hotspots threatening CES supply, using the five standardised constraining features. All features were considered equally important, covering complementary aspects of CES supply, and were equally weighted. The resulting hotspots of CES attractions and constraints were classified into four main classes – low, medium, high and very high – using the Jenks Natural Breaks in ArcMap 10.1 (ESRI, 2012). This approach

150 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

provides natural groupings inherent in the data, by identifying groups of similar attractions and constraints and maximising differences between classes (Casado-Arzuaga et al., 2014; Peña et al., 2015). To identify the areas where many attractions are threatened or not by many constraints, the distributions of classified attraction and constraint hotspots were overlapped to generate the combined attraction-constraint hotspots (as in Santarém et al., 2019a). The hotspots combining many attractions with few constraints overlapped the class very high attractions with the class low constraints, while the hotspots combining both many attractions and many constraints overlapped the class very high attractions with the classes high and very high constraints.

Results

Distribution of attraction and constraint features

Attraction features The Sahara-Sahel displays substantial spatial heterogeneity in the distribution of attraction features related with CES supply (Fig. 3.9). Species richness is higher in the Sahel in comparison to the Sahara, a pattern that contrasts with the higher number of endemics and flagships found in the Sahara along the Western and Eastern corridors, mountains, and wetlands. Ecoregions are heterogeneously distributed along the region, but parts of Mauritania, Algeria and Egypt display contact zones between multiple ecoregions. Forest reserves are dense along the Senegal River and in southern Niger. Protected areas are generally widespread, but some regions display denser levels, such as along the Morocco-Algeria, Mauritania-Senegal, and Algeria-Libya borders, in central Tunisia, and Lake Chad surroundings. Some countries (e.g., Egypt) protect much more parcels of land than others (e.g., Libya). World Heritage Sites are heterogeneously spread across the region, but denser levels can be found along the Algeria-Libya border. There is no clear geographic structure in the distribution of all landscape features, but some regions tend to concentrate higher levels, such as central Mauritania, southern Algeria, and along the Egypt-Libya-Sudan borders. Gorges and mountain passes are well distributed in Sahara-Sahel, but higher concentrations can be found mainly along Morocco-Algeria border. Rock formations are numerous in Mauritania and Chad, but the highest density can be found in only one pixel in southern Niger. High density of caves can be found in north-eastern Libya. Wetlands are denser in the wettest parts of the Sahel, along the Niger and Senegal rivers, and Lake Chad. Rock-pools (gueltas) are concentrated in the Saharan mountains. Ethnic groups richness is higher along FCUP 151 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Sahel parts of Mauritania, Mali and Chad. Oases concentrate particularly in central Niger, and along the Algeria-Tunisia and Eritrea-Ethiopia-Sudan borders. Monuments are overall rare in Sahara-Sahel and can be found mostly in Egypt and Algeria. Caravan villages are spread across the study area, as well as caravan routes. Fortifications concentrate mostly in Algeria and Morocco. Ruins and sites of historical occupation are mostly found in northern Sahara (Morocco, Tunisia, and northern Libya) and along the Nile River. Rock art is denser in southern Algeria and Libya, along the Libya-Egypt-Sudan borders, and in Egypt.

152 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 3.9. Distribution of standardised attractions to cultural ecosystem services supply in the Sahara-Sahel at 0.5-degree resolution. See Table 3.3 for codes of variable.

FCUP 153 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Constraint features Considerable spatial heterogeneity in the distribution of constraint features related with CES supply was found (Fig. 3.10). Landmines are denser along the Morocco-Mauritania border, in northern Chad, and along the Libya-Chad and Eritrea-Sudan borders. Attacks and violence are denser in northern Egypt and Sudan. Smuggling and human migration routes are spread in Sahara-Sahel but tend to concentrate along borders between countries. Exploitation of natural resources is highly concentrated in north-eastern Algeria and northern Libya.

Fig. 3.10. Distribution of standardised constraints to cultural ecosystem services supply in the Sahara-Sahel at 0.5-degree resolution. See Table 3.3 for codes of variable.

Hotspots of cultural ecosystem services attractions and constraints The spatial distribution of hotspots of attractions and constraints for CES supply in Sahara- Sahel varied substantially (Fig. 3.11). Large areas concentrate the highest level of attractions, especially in most of the main Sahara-Sahel mountains, along hydrographic networks (Senegal, Niger and Nile rivers, Lake Chad, and Chott El Jerid), and in south-eastern Morocco. Isolated

154 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

attraction hotspots are also found, for instance in Egypt, Algeria, Chad or coastal Mauritania. Large flat areas tend to concentrate fewer attractions. Overall, 35.4% of the study area displays high (27.9%) to very high (7.5%) levels of attractions to CES supply. High to very high levels of constraints can be found along the Morocco-Mauritania and Algeria-Libya borders, north-eastern Algeria, Mali, Libya, northern Egypt, and western and eastern Sudan. Areas exhibiting low levels of constraints are mostly coincident with the areas displaying low levels of attractions. Overall, 8.6% of the study area displays high (7.5%) to very high (1.1%) levels of constraints impacting CES supply. The combined distributions of attractions and constraints to CES supply also varied substantially (Fig. 3.11). Contiguous pixels displaying very high levels of attractions and very few constraints are found along central and southwestern Mauritania, Nile and Senegal rivers, Chad (Ennedi mountain), and southeast Morocco, while isolated pixels can be found in south-eastern Egypt and Mauritania, and central Chad. Furthermore, contiguous pixels displaying both very high levels of attractions and constraints are found in Algeria (Tassili-n’Ajjer and Hoggar mountains), Niger (Aïr mountains), Lake Chad, Niger River, and lower Nile River in north-eastern Egypt, while isolated pixels are found in eastern Sudan, north-eastern Libya and Mali, and in the Algeria- Tunisia-Libya border. FCUP 155 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 3.11. Distribution of hotspots of attraction and constraint features to cultural ecosystem services supply in the Sahara-Sahel at 0.5-degree resolution (top and middle) and of hotspots combining many attractions with few constraints and simultaneously many attractions and constraints (bottom).

Discussion

To our best knowledge, this is the first study mapping and analysing transboundary constraints threatening the supply of CES at the sub-continental scale using primary and

156 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

secondary data. This is one of the first studies mapping and analysing ecosystem services in the Sahara-Sahel region and provides the largest and the most comprehensive database on the natural and historical-cultural heritage of the region to date. Integrating attractions and constraints for mapping and analysing CES supply in the largest warm region of the world, as we did here, provided a tool that allows policymakers to identify priority areas of intervention in order to promote sustainable use of the arid ecosystems. Next, we discuss some spatial patterns found, as well as possible management applications, and limitations of the work that deserve future improvement.

Spatial patterns of CES supply in the Sahara-Sahel Overall, 35.4% of the Sahara-Sahel area seems to supply high to very high levels of CES, a result similar to what was found in other continental-scale (e.g., 38% in Europe; Parachinni et al., 2014) and local-scales studies (e.g. 44% in Basque Country, Spain; Peña et al., 2015). This finding suggests that despite the generalised lack of research interest in studying desert ecosystem services (Lu et al., 2018), the biome displays high recreational values that need to be further explored. Recent global valuations of ecosystem services demonstrated that deserts are among the biomes providing the highest monetary value per area (Costanza et al., 2014; Kubiszewski et al., 2017). If the monetary value of desert CES remains to be calculated (Davies et al., 2012), our findings suggest that this figure can be high in comparison to other biomes. Additionally, although remoteness and harsh environments characterise deserts, the sparsely distributed vegetation and uncovered large landscapes facilitates the observation of large-bodied species (Safriel and Adeel, 2005; Santarém et al., 2019a,b) and, therefore, these species may supply more CES in deserts than in any other biome where vegetation may limit their sight. Hence, adopting a “Green Economic Growth” strategy in deserts will help protecting natural and cultural assets and maintaining the supply of ecosystem services on which human development and well- being depend on (Safriel and Adeel, 2005; Davies et al., 2012; Bidak et al., 2015; Ghazi et al., 2018). As initially hypothesised, most of the attraction features mapped tended to concentrate in or around the main mountains and wetlands (Fig. 3.9; Fig. C1), which linked them with hotspots of CES supply (Fig. 3.11). Mountains and wetlands have also been identified as hotspots in other continental-scale (Paracchini et al., 2014; van Zanten et al., 2016; Komossa et al., 2018) and local-scale analyses (Plieninger et al., 2013; Peña et al., 2015). As wetlands tend to concentrate higher levels of attraction features, ecotourists probably tend to attribute an increased importance to wetlands availability (Scholte et al., 2018), especially when mountains and wetlands are protected (Paracchini et al., 2014; van Zanten et al., 2016) and where cultural heritage is dense FCUP 157 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

(van Berkel and Verburg, 2014; Komossa et al., 2018). Despite Sahara-Sahel mountains and wetlands remain largely under-sampled, they constitute refugia for local biodiversity (Brito et al., 2014; Brito and Pleguezuelos, 2019) and support key ecosystem services (Davies et al., 2012), and were here revealed as main CES suppliers. Corridors along the Nile River and the Atlantic Sahara also showed a strong CES supply, reinforcing their key role in preserving the desert natural capital (Brito et al., 2014). The flat and sandy parts of the Sahara-Sahel apparently displayed fewer attractions for CES supply. These areas tend to be highly homogeneous and are thus less demanded by ecotourists for recreation (Peña et al., 2015). Yet, as vast empty-quarters and dune massifs are crucial for the persistence of large flagship Sahara species, such as the Addax (Addax nasomaculatus) and the Scimitar-horned oryx (Oryx dammah) (Durant et al., 2014; Brito et al., 2018; Santarém et al., 2019a), they may supply moderate to high levels of other ecosystem services at finer scales (e.g., provisioning services; Wei et al., 2018). Future studies could explore the validity of this assumption. Hotspots of constraint features matched our initial hypothesis that they would tend to occur in areas where human-led threats were preliminarily identified (see Brito et al., 2018). For instance, the current conflict, the high density of natural resources exploration activities, and the smuggling routes converging in northern Libya suggest the region as no-go area for ecotourists. Additionally, in areas where the number of conflicts is high (e.g., southern Sudan or northern Egypt), our study revealed hotspots of constraints that likely alter local ecosystems. For instance, many World Heritage Sites were identified as severely threatened by terrorist group activities. Thus, it is urgent the deliberation of crucial avenues for global transparent and real-time mechanisms to tackle the risks of permanent loss of these sites of outstanding importance to humanity (Levin et al., 2019). Some of the areas concentrating very high levels of attractions spatially matched with areas displaying very high levels of constraints, such as the cases of the Niger River valley and the Lake Chad surroundings. Despite being among the places displaying more attractions for CES supply, they are downgraded by pixels exhibiting multiple constraints threatening CES supply. Similar cases occur in isolated clusters, for instance in the lower Nile valley, where the number of attacks against civilians and migration routes are likely impacting local CES supply. This overlap between areas of high CES supply and human threats exacerbate the challenges to preserve the natural and cultural heritage of Sahara-Sahel and emphasise the need to carefully consider human-wildlife mitigation measures (Brito et al., 2014, 2016, 2018).

158 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Large clusters displaying very high levels of attractions with very few, if any, constraints were identified, for instance in Morocco, Mauritania, Algeria, Niger and Chad. These areas allow maximising the returns of visiting time, i.e., where ecotourists can observe different attractions in the same region without the need to travel between distant regions, and with minimum jeopardy from constraint features. Still, our results should be interpreted cautiously, as areas currently depicting high levels of attractions or constraints may alter in the future. Yet, this CES mapping exercise provides ecotourists, planners, and companies with a preliminary view, though not optimized (see section 4.2.), on which are the safest areas to visit in the Sahara-Sahel.

Management of CES in Sahara-Sahel The combined map of attractions and constraints hotspots in Sahara-Sahel revealed considerable areas supplying CES. Yet, recreational use in a form of ecotourism in Sahara-Sahel is currently very asymmetrical and not all the attractions are equally harnessed. For example, UNESCO World Heritage Sites and oases are being explored by local tour companies that offer cultural experiences (Santarém et al., 2019b), but most of these experiences are located for instance in Morocco and Algeria, while Mauritania and Chad remain poorly visited (UNWTO, 2018b). Here, we identified other areas and values that could be also explored and that could improve local economies where well-known international cultural attractions are lacking. For example, southern Mauritania displays one of the largest populations of the West-African crocodile (Crocodylus suchus) inhabiting the desert biome, a flagship species whose observation through detailed flagship-based ecotourism programmes could generate additional income sources to local people (Brito et al., 2014). Direct and indirect jobs in the ecotourism sector and the improvement of physical infrastructures for education (schools) and aid assistance (hospitals, clinics) can generate an extra incentive to protect local biodiversity. Additionally, diverting from cultural-only tourism to ecotourism can also contribute to the long-term preservation of deserts biomes, as environmental awareness among ecotourists and locals improves, and sustainable strategies for local development start to be a priority in policymaking (UNEP, 2006b; Santarém et al., 2019b). Still, there are several aspects that need improvement, such as the limiting entry requirements (VISA politics) of several countries or the "negative" impression given by past conflicts that detracts international ecotourists from travelling to North-Africa. Bureaucracy needs to be facilitated to improve confidence and specialised ecotourism requests need to be further explored. Our study identified many transboundary CES hotspots, which helped understanding the diversity and similarity of continent’s cultural-historical backgrounds and emphasised the need to FCUP 159 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel map and analyse CES across borders (van Zanten et al., 2016). For instance, while the utilization of CES along the Algeria-Niger border is facilitated, it is impossible along the Morocco-Mauritania border due to the heavily mined 2,700km military berm dividing these countries, which detracts any form of travelling. Additionally, mapping and analysing CES transboundary helped to identify priorities for ecosystem management and restoration across borders (Naidoo et al., 2008; Schulp et al., 2014). For instance, the populations of the African savannah elephant (Loxodonta africana) can be observed in Mali during the dry season, but during the wet season ecotourists need to travel to Burkina Faso in order to follow their movements (Wall et al., 2013). Similarly, a deep understanding of how past societies have lived in the region is only possible if ecotourists visit both the mountains of Tassili-n’Ajjer in Algeria and Tadrart Acacus in Libya, which are considered a “cultural province” (Gallinaro, 2013). In this paper we have stressed the importance of considering the preservation of natural and cultural assets through the development of transboundary "green" infrastructure networks that benefit neighbouring societies (Naidoo et al., 2008; Schulp et al., 2014) and of including indigenous communities in regional plans aiming to maximise CES supply. By mapping ecosystem services transboundary, decision-makers may identify areas where changes could impact ecosystems, classify avoidance-areas for ecotourists and allocate recreation resources thoroughly (Lanouar and Goaied, 2019). Despite the high concentration of constraints in some regions, a differentiation between their intrinsic characteristics and consequences for wildlife and ecotourists should be done. On the one hand, there are constraints that impact mostly biodiversity features, such as oil and mining facilities that promote the extirpation of charismatic megafauna (Duncan et al., 2014; Brito et al., 2018). These constraints diminish the possibilities of ecotourists to observe Sahara-Sahel charismatic species. On the other hand, there are constraints that may be capital to people’s life, such as landmines (Brito et al., 2018; Dioko and Harrill, 2019), and that may clearly detract ecotourists from visiting certain areas of the Sahara-Sahel. Measuring the time length of each constraint would be pertinent for policymakers prevent socio-economic costs related to the loss of desert CES and, consequently, from ecotourists to visit certain areas. Doing so will help determining which type of policies should be implemented to recover and enhance ecosystems (Lanouar and Goaied, 2019). Still, local and international decision-makers should reinforce transboundary strategic actions to manage biodiversity and the ecosystem services derived by it to halt the ongoing destruction of natural resources. Overall, ensuring Sahara-Sahel CES supply will rely on: 1) detailed mapping of conflict hotspots; 2) increasing the levels of international and regional investment in human development, nature preservation, and technology transfer; 3)

160 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

reinforcing policies to curb ecosystem degradation and to revert prejudiced effects of war, landmines and kidnapping on local ecotourism; 4) developing community-based natural resources management policies; and 5) capitalising local traditional knowledge (Safriel and Adeel, 2005; Egoh et al., 2007; Brito et al., 2014, 2018; Santarém et al., 2019b). We found that many of the regions responsible for high levels of CES supply are insufficiently covered by the current network of protected areas (Fig. 3.9). For instance, Mauritania displays many attraction hotspots (Fig. 3.11) but landscapes, species and ecosystems are poorly protected (IUCN and UNEP-WCMC, 2017). We sound the urgency to develop protecting schemes towards this unique desert biodiversity that deliver key ecosystem services (Brito et al., 2016; Santarém et al., 2019a). Ecological corridors and transboundary mega conservation areas should be prioritised to preserve these services (Davies et al., 2012; Egoh et al., 2012; Brito et al., 2016).

Limitations and further research This work displays some limitations that may have locally biased our results. For instance, species distributions were based in IUCN polygons depicting extents of occurrence, which are often criticised because they may overestimate species’ ranges (Graham and Hijmans, 2006; Chung et al., 2018), or in the case of Sahara-Sahel, they may underestimate distributions due to scarce sampling efforts (Brito et al., 2014, 2016). This may have probably impacted the accuracy of the species richness maps. Similar issues may arise from other world databases used to derive variables useful for assessing ecosystem services (Pandeya et al., 2016). Additionally, some regions may be supplying high to very high levels of attractions, but the generalised paucity of data dictated a different scenario in our study. For instance, the Tibesti Mountain (Chad) is the highest Saharan mountain and displays several landscape features of outstanding value (e.g., volcanic cones and meteor craters; Santarém et al., 2019b); yet, it is among the regions with the least available high-resolution maps of natural features (Brito et al., 2014). Although these constraints may bias the identified CES hotspots, their effects were likely diluted using a coarse spatial resolution of 0.5 degrees (Weyland and Laterra, 2014). Future studies should make use of accurate distribution data to overcome this issue whenever possible, for example by deriving species ranges from ecological niche-based models, as suggested by Santarém et al. (2019a). Using a coarse spatial resolution may compromise the efficiency in identifying areas supplying CES. Yet, such resolution is required when the large extent of the study area and the computational power needed to perform the analysis demands such compromise (Brito et al., 2016). Even if the spatial resolution here used constrained the real CES measures at local scales, this study offered hints to framework regional CES planning that can be applied at finer-scale FCUP 161 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel resolutions. Social media opens new avenues for data mining to locate CES in the space and time. Social networks containing geo-tagged data allow collecting information about CES at the fastest pace ever (Oteros-Rozas et al., 2018; Vaz et al., 2019). This new trend to gather CES data will be useful only if, and when, ecotourism increases in Sahara-Sahel to levels that will allow assembly such data. The spatial and temporal dynamics of attractions and constraints in Sahara-Sahel were not here specifically considered. For example, the salt-caravans crossing the Ténéré (Niger) are only observable in certain periods of the year, some wetlands are seasonal and only available during the rainy season, and wintering birds or reptiles in northern latitudes are only observable during specific periods of the year. Constraints to CES supply may also be highly dynamic. For instance, long-term regional conflict and lack of formal land access points virtually closed Mauritania to international travelling until the early 2000’s, which changed afterwards with the amelioration of security conditions and the opening of an official border post, and again changed after 2008 following localised terrorist attacks throughout the country, to change again around 2018 with increasing of security conditions. The dynamics of these processes are very fast and thus hard to include in general assessments of CES supply. Still, researchers are requesting to contemplate interlinked spatial-temporal dynamics of CES (e.g., Rieb et al., 2017; Small et al., 2017) because they may impact patterns and processes of CES and drive better long-term outcomes if considered. But, in regions like Sahara-Sahel, where scientific data is scarce and spatially fragmented (see Brito et al., 2014), mapping biophysical provision of CES will be constrained until accurate data is available (Small et al., 2017). The maps here produced only provide possible paths for sustainable desert ecotourism development. Scholars are requesting that studies contemplate not only the supply side, but also the demand side of CES (Wolff et al., 2015). This may also partially challenge some of the supply side results or their implementation as individual ecotourists can experience and value attractions and CES differently, i.e., “there is no consensus regarding what constitutes ‘value’ of nature for individuals” (Stålhammar and Pedersen, 2017, p. 2). Still, optimized solutions could be derived for cultural tourism, landscape tourism and ecotourism, for example, tailored according to the preferences of distinct visiting groups. Additionally, the different impacts of constraints on CES supply (or any other services such as provisioning, regulating, and supporting; MEA, 2005) can be explored in future studies by attributing different weights to the constraints that are capital to human life (e.g., landmines) or that diminish local attractions (e.g., gas pipelines). Spatial decision-support tools present an excellent tool to further explore this, as researchers can attribute different weights to variables according to the objective of the work (Moilanen et al.,

162 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

2014). They have been used to study how multiple ecosystem services are integrated into conservation planning at regional geographical scales (Chan et al., 2006; Hermoso et al., 2018) but conservation planning exercises applied to CES supply at continental scales remain to be done. The usefulness of using these tools at continental scales has been proved, for instance, in identifying priority conservation areas in Sahara-Sahel (Brito et al., 2016), but future studies could use them also to establish spatial priorities for developing recreation and ecotourism for different society segments while accounting for factors constraining CES supply.

Conclusions

Evaluating hotspots of attractions and constraints to CES supply in conflict regions is an important step to counteract ecosystem degradation and develop tools to improve their provision. The approach developed here is scalable and replicable worldwide and the criteria used could set the guidelines to identify the regions supplying the highest levels of CES in data-scarce regions. We highlight the significance of using raw data to robustly identify the areas supplying CES that are vulnerable to human-mediated constraints. Including conflicts on ecosystem services research as we did here is important to develop the field of research even further. We also note the importance of considering the human, social and natural capital (people, cultural societies, and the environment, respectively) to map the benefits that ecosystems provide to people. This brings additional importance in the case of low-income countries and poor peripheral regions where conservation and ecotourism efforts need to be reinforced in order to protect the environment and the local people depending on it for sustainable development. Ecosystem services planning should also involve multidisciplinary and transdisciplinary teams of scholars and practitioners to fully integrate theoretical and empirical expertise from diverse fields of knowledge – biology, geosciences, geography, economics, and social sciences – and to guide conservation management efforts efficiently (Chan et al., 2006; Naidoo et al., 2008; Rosa et al., 2017). It is by integrating multidisciplinary teams that scholars can maximise the benefits of CES, critical to poverty alleviation (UN-SDGs 1 and 8) and to biodiversity conservation (UN-SDG 15), even in regions under geopolitical conflicts that press nature preservation. In future, integrating science and policy will be critical to sustain ecosystem services and the natural capital (Burkhard et al., 2012b), and policy makers can use the provided framework to help achieve regional targets for UN-SDGs.

FCUP 163 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

References

Balmford, A., Green, J.M.H., Anderson, M., Beresford, J., Huang, C., Naidoo, R., Walpole, M., Manica, A. (2015) Walk on the wild side: Estimating the global magnitude of visits to protected areas. PLoS Biology, 13(2): e1002074. doi: https://doi.org/10.1371/journal.pbio.1002074. Barnett, T., Guagnin, M. (2014). Changing places: rock art and Holocene landscapes in the Wadi al-Ajal, south-west Libya. Journal of African Archaeology, 12: 165-182. doi: https://doi.org/10.3213/2191-5784-10258. Beauchamp, Z. (2014a). Libya's horrible, chaotic year, in one map. Vox World. Accessed at www.vox.com/2014/12/25/7447099/libya-conflict-map. 2014. Beauchamp, Z. (2014b). The crisis in Nigeria, in 11 maps and charts. Vox World. Accessed at www.vox.com/2014/5/13/5710484/boko-haram-maps-charts-nigeria. Biagetti, S., Cancellieri, E., Cremaschi, M., Gauthier, C., Gauthier, Y., Zerboni, A., Gallinaro, M. (2013). The ‘Messak Project’: Archaeological research for cultural heritage management in SW Libya. Journal of African Archaeology, 11: 55-74. doi: 10.3213/2191-5784-10231. Bidak, L.M., Kamal, S.A., Halmy, M.W.A., Heneidy, S.Z. (2015). Goods and services provided by native plants in desert ecosystems: Examples from the northwestern coastal desert of Egypt. Global Ecology and Conservation, 3: 433-447. doi: http://dx.doi.org/10.1016/j.gecco.2015.02.001.Brachet, J., Choplin, A., Pliez, O. (2011). Le Sahara entre espace de circulation et frontière migratoire de l'Europe. Herodote, 142: 163- 182. doi: https://doi.org/10.3917/her.142.0163. Brémont, A. (2018). Des éléphants, des hippopotames et des mouflons. Trois hypothèses de marqueurs animaux d’identités régionales pendant les périodes de Nagada I-II. Archéo-Nil 28: 69-98. Brito, J.C., Durant, S.N., Pettorelli, N., Newby, J., Canney, S., Algadafi, W., Rabeil, T., Crochet, P.-A., Pleguezuelos, J.M., Wacher, T., de Smet, K., Gonçalves, D.V., Ferreira da Silva, M.J., Martínez-Freiría, F., Abáigar, T., Campos, J.C., Comizzoli, P., Fahd, S., Fellous, A., Malam Garba, H.H., Hamidou, D., Harouna, A., Hatcha, M.S., Nagy, A., Silva, T.L., Sow, A.S., Vale, C.G., Boratyński, Z., Rebelo, H., Carvalho, S.B. (2018). Armed conflicts and wildlife decline: challenges and recommendations for effective conservation policy in the Sahara-Sahel. Conservation Letters, 11: e12446. doi: https://doi.org/10.1111/conl.12446. Brito, J.C., Godinho, R., Martínez-Freiría, F., Pleguezuelos, J.M., Rebelo, H., Santos, X., Vale, C.G., Velo-Antón, G., Boratyński, Z., Carvalho, S.B., Ferreira, S., Gonçalves, D.V., Silva, T.L.,

164 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Tarroso, P., Campos, J.C., Leite, J.V., Nogueira, J., Álvares, F., Sillero, N., Sow, A.S., Fahd, S., Crochet, P.-A., Carranza, S. (2014). Unravelling biodiversity, evolution and threats to conservation in the Sahara-Sahel. Biological Reviews, 89: 215-231. doi: https://doi.org/10.1111/brv.12049. Brito J.C., Pleguezuelos J.M. (2019). Desert Biodiversity—World’s Hot Spots/Globally Outstanding Biodiverse Deserts. In: Encyclopedia of the World’s Biomes. doi: https://doi.org/10.1016/B978-0-12-409548-9.11794-4. Brito, J.C., Tarroso, P., Vale, C.G., Martínez-Freiría, F., Boratyński, Z., Campos, J.C., Ferreira, S., Godinho, R., Gonçalves, D.V., Leite, J.V., Lima, V.O., Pereira, P., Santos, X., Ferreira da Silva, M.J., Silva, T.L., Velo-Antón, G., Veríssimo, J., Crochet, P.-A., Pleguezuelos, J.M., Carvalho, S.B. (2016). Conservation biogeography of the Sahara-Sahel: additional protected areas are needed to secure unique biodiversity. Diversity and Distributions, 22: 371-384. doi: https://doi.org/10.1111/ddi.12416. Brown, G., Fagerholm, N. (2015) Empiritical PPGIS/PGIS mapping of ecosystem services: A review and evaluation. Ecosystem Services, 13: 119-133. doi: https://doi.org/10.1016/j.ecoser.2014.10.007. Burkhard, B., Kroll, F., Nedkov, S., Müller, F. (2012a). Mapping ecosystem service supply, demand and budgets. Ecological Indicators, 21: 17–29. doi: 10.1016/j.ecolind.2011.06.019. Burkhard, B., de Groot, R., Costanza, R., Seppelt, R., Jørgensen, S.E., Potschin, M. (2012b). Solutions for sustaining natural capital and ecosystem services. Ecological Indicators, 21: 1– 6. doi: 10.1016/j.ecolind.2012.03.008. Casado-Arzuaga, I., Onaindia, M., Madariaga, I., Verburg, P. H. (2014). Mapping recreation and aesthetic value of ecosystems in the Bilbao Metropolitan Greenbelt (northern Spain) to support landscape planning. Landscape Ecology, 29(8): 1393–1405. doi: https://doi.org/10.1007/s10980-013-9945-2. Cerretelli, S., Poggio, L., Gimona, A., Yakob, G., Boke, S., Habte, M., Coull, M., Peressotti, A., Black, H. (2018). Spatial assessment of land degradation through key ecosystem services: The role of globally available data. Science of The Total Environment, 628-629: 539–555. doi: 10.1016/j.scitotenv.2018.02.085. Chan, K.M.A., Shaw, M.R., Cameron, D.R., Underwood, E.C., Daily, G.C. (2006). Conservation Planning for Ecosystem Services. PLoS Biology, 4(11): e379. doi: 10.1371/journal.pbio.0040379. FCUP 165 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Cooper, A., Shine, T., McCann, T., Tidane, D.A. (2006). An ecological basis for sustainable land use of Eastern Mauritanian wetlands. Journal of Arid Environments, 67: 116–141. doi: https://doi.org/10.1016/j.jaridenv.2006.02.003. Chung, M.G., Dietz, T., Liu, J., (2018). Global relationships between biodiversity and nature- based tourism in protected areas. Ecosystem Services, 34: 11-23. doi: http://doi.org/10.1016/j.ecoser.2018.09.004. Costanza, R., de Groot, R., Sutton, P., van der Ploeg, S., Anderson, S.J., Kubiszewski, I., Farber, S., Turner, R.K. (2014). Changes in the global value of ecosystem services. Global Environmental Change, 26: 152–158. doi: 10.1016/j.gloenvcha.2014.04.002. Crossman, N.D., Burkhard, B., Nedkov, S., Willemen, L., Petz, K., Palomo, I., Drakou, E.G., Martín-López, B., McPhearson, T., Boyanova, K., Alkemade, R., Egoh, B., Dunbar, M.B., Maes, J. (2013). A blueprint for mapping and modelling ecosystem services. Ecosystem Services, 4: 4–14. doi: 10.1016/j.ecoser.2013.02.001. Daniel, T.C., Muhar, A., Arnberger, A., Aznar, O., Boyd, J.W. Chan, K.M.A., Costanza, R., Elmqvist, T., Flint, C.G., Gobster, P.H., Grêt-Regamey, A., Lave, R., Muhar, S., Penker, M., Ribe, R.G., Schauppenlehner, T., Sikor, T., Soloviy, I., Spierenburg, M., Taczanowska, K., Tam, J., von der Dunk, A. (2012). Contributions of cultural services to the ecosystem services agenda. PNAS, 109(23): 8812-8819. Davies, J., Poulsen, L., Schulte-Herbrüggen, B., Mackinnon, K., Crawhall, N., Henwood, W.D., Dudley, N., Smith, J., Gudka, M. (2012). Conserving Dryland Biodiveristy. IUCN: Nairobi, . de Hass, H. (2007). The myth of invasion. Irregular migration from West Africa to the Maghreb and the European Union. International Migration Institute Research Report, 79 pp. Accessed at http://www.heindehaas.com/Publications/de%20Haas%202007%20Irregular%20migration%2 0from%20West%20Africa%20.pdf. DesertInfo (2006). Landmines: pictures, maps, GPS + Google Earth Data. DesertInfo Forum. Accessed on 06 September 2018 at http://www.desert-info.ch/desert-info- forum/viewtopic.php?t=1927. Díaz, S., Settele, J., Brondízio, E. (2019). Summary for policymakers of the global assessment report on biodiversity and ecosystem services of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services – ADVANCE UNEDITED VERSION. Accessed on 28 May 2019 at https://www.ipbes.net/news/ipbes-global-assessment-preview.

166 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Dinerstein, E., Olson, D., Joshi, A., Vynne, C., Burgess, N.D., Wikramanayake, E., Hahn, N., Palminteri, S., Hedao, P., Noss, R., Hansen, M., Locke, H., Ellis, E.C., Jones, B., Barber, C.V., Hayes, R., Kormos, C., Martin, V., Crist, E., Sechrest, W., Price, L., Baillie, J.E.M., Weeden, D., Suckling, K., Davis, C., Sizer, N., Moore, R., Thau, D., Birch, T., Potapov, P., Turubanova, S., Tyukavina, A., de Souza, N., Pintea, L., Brito, J.C., Llewellyn, O.A., Miller, A.G., Patzelt, A., Ghazanfar, S.A., Timberlake, J., Klöser, H., Shennan-Farpon, Y., Kindt, R., Lillesø, J.-P.B., van Breugel, P., Graudal, L., Voge, M., Al-Shammari, K.F. and Saleem, M. (2017). An ecoregion-based approach to protecting half the terrestrial realm. BioScience, 67: 534-545. doi: 10.1093/biosci/bix014. Dioko, L.(Don)A.N., Harrill, R. (2019). Killed while traveling – Trends in tourism-related mortality, injuries, and leading causes of tourist deaths from published English news reports, 2000–2017 (1H). Tourism Management, 70: 103–123. doi: 10.1016/j.tourman.2018.08.002. Duncan, C., Kretz, D., Wegmann, M., Rabeil, T., Pettorelli, N. (2014). Oil in the Sahara: mapping anthropogenic threats to Saharan biodiversity from space. Philosophical Transactions of the Royal Society of London B., 369: 20130191. doi: 10.1098/rstb.2013.0191. Durant, S.M., Pettorelli, N., Bashir, S., Woodroffe, R., Wacher, T., De Ornellas, P., Ransom, C., Abáigar, T., Abdelgadir, M., el Alqamy, H., Beddiaf, M., Belbachir, F., Belbachir-Bazi, A., Berbash, A.A., Beudels-Jamar, R., Boitani, L., Breitenmoser, C., Cano, M., Chardonnet, P., Collen, B., Cornforth, W.A., Cuzin, F., Gerngross, P., Haddane, B., Hadjeloum, M., Jacobson, A., Jebali, A., Lamarque, F., Mallon, D., Minkowski, K., Monfort, S., Ndoassal, B., Newby, J., Ngakoutou, B.E., Niagate, B., Purchase, G., Samaïla, S., Samna, A.K., Sillero-Zubiri, C., Soultan, A.E., Stanley Price, M.R., Baillie, J.E.M. (2012). Forgotten Biodiversity in Desert Ecosystems. Science, 336(6087): 1379–1380. doi: 10.1126/science.336.6087.1379. Durant, S.M., Wacher, T., Bashir, S., Woodroffe, R., De Ornellas, P., Ransom, C., Newby, J., Abáigar, T., Abdelgadir, M., el Alqamy, H., Baillie, J., Beddiaf, M., Belbachir, F., Belbachir- Bazi, A., Beudels-Jamar, R., Boitani, L., Breitenmoser, C., Cano, M., Chardonnet, P., Collen, B., Cornforth, W.A., Cuzin, F., Gerngross, P., Haddane, B., Hadjeloum, M., Jacobson, A., Jebali, A., Lamarque, F., Mallon, D., Minkowski, K., Monfort, S., Ndoassal, B., Niagate, B., Purchase, G., Samaïla, S., Samna, A.K., Sillero-Zubiri, C., Soultan, A.E., Stanley Price, M.R., Pettorelli, N. (2014). Fiddling in biodiversity hotspots while deserts burn? Collapse of the Sahara’s megafauna. Diversity and Distributions, 20: 114–122. doi: 10.1111/ddi.12157. Egoh, B., Rouget, M., Reyers, B., Knight, A.T., Cowling, R.M., van Jaarsveld, A.S., Welz, A. (2007) Integrating ecosystem services into conservation assessments: A review. Ecological Economics, 63: 714.721. doi: https://doi.org/10.1016/j.ecolecon.2007.04.007. FCUP 167 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Eigenbrod, F., Armsworth, P.R., Anderson, B.J., Heinemeyer, A., Gillings, S., Roy, D.B., Thomas, C.D., Gaston, K. J. (2010). The impact of proxy-based methods on mapping the distribution of ecosystem services. Journal of Applied Ecology, 47(2): 377–385. doi: 10.1111/j.1365- 2664.2010.01777.x. Environmental Systems Research Institute (ESRI). (2012). ArcGIS Release 10.1. Redlands, CA, USA. Ewi. (2010). A decade of kidnappings and terrorism in West Africa and the Trans-Sahel region. African Security Review, 19: 64-71. doi: 10.1080/10246029.2010.539812. Fennell, D. (2015). Ecotourism (4th ed.). Abingdon, UK: Routledge. Gallinaro, M. (2013). Saharan rock art: local dynamics and wider perspectives. Arts, 2: 350-382. doi: https://doi.org/10.3390/arts2040350. Gauthier, Y., Gauthier, C. (2006). Nouveaux abris peints de l’Ennedi (Chad). Sahara, 17: 165- 172. Gauthier, Y., Gauthier, C. (2008). Art rupestre, monuments funéraires et aires culturelles: Nouveaux documents concernant le Messak, le Sud-Est du Fezzân et l’Oued Djerat. Cahiers de l’ARRS, 12: 89-104. Gauthier, Y., Gauthier, C. (2011). Les lacs de Têh-n-Beka: Contribution des gravures à la connaissance du climat à l’Holocène. Les Cahiers de l’AARS, 15: 47-86. GeoCover (2018). Landsat GeoCover. Global Land Cover Facility. Accessed on 06 September 2018 at http://www.landcover.org/research/portal/geocover/. Ghazi, H., Messouli, M., Yacoubi Khebiza, M., Egoh, B.N. (2018). Mapping regulating services in Marrakesh Safi region - Morocco. Journal of Arid Environments. 159: 54-65. doi: 10.1016/j.jaridenv.2018.03.005. Gigović, L., Pamučar, D., Lukić, D., Marković, S. (2016). GIS-Fuzzy DEMATEL MCDA model for the evaluation of the sites for ecotourism development: A case study of “Dunavski ključ” region, Serbia. Land Use Policy, 58: 348-365. doi: 10.1016/j.landusepol.2016.07.030. Graham, C.H., Hijmans, R.J. (2006). A comparison of methods for mapping species ranges and species richness. Global Ecology and Biogeography, 15(6): 578-587. doi: 10.1111/j.1466- 822x.2006.00257.x. Grossman, L. (2015). Boko Haram continues to slaughter Nigerians. The Long War Journal. Accessed at www.longwarjournal.org/archives/2015/01/boko_haram_attacks_c.php. Hall, C.M. and Saarinen, J. (2010). Last Chance to See? Future Issues for Polar Tourism and Change. In Hall, C.M & J Saarinen (eds), Tourism and Change in the Polar Regions: Climate, Environment and experiences (pp. 301-310). London: Routledge.

168 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Hanaček, K., Rodríguez-Labajos, B. (2018). Impacts of land-use and management changes on cultural agroecosystem services and environmental conflicts—A global review. Global Environmental Change, 50: 41–59. doi: 10.1016/j.gloenvcha.2018.02.016. Harmon, S.A. (2016). Terror and Insurgency in the Sahara-Sahel Region: Corruption, Contraband, Jihad and the Mali War of 2012-2013. London, UK: Routledge. Hermoso, V., Cattarino, L., Linke, S., Kennard, M.J. (2018). Catchment zoning to enhance co- benefits and minimize trade-offs between ecosystem services and freshwater biodiversity conservation. Aquatic Conservation: Marine and Freshwater Ecosystems, 28: 1004-1014. doi: 10.1002/aqc.2891. IUCN (2018). The IUCN Red List of Threatened Species, v 6.1. International Union for Conservation of Nature and Natural Resources. Accessed on 14 November 2018 at http://www.iucnredlist.org. IUCN, UNEP-WCMC (2018). The World Database on Protected Areas (WDPA). Cambridge, UK: UNEP-WCMC. Accessed on 15 November of 2018 at http://www.protectedplanet.net. Jesse, F., Keding, B., Pöllath, N., Bechhaus-Gerst, M. and Lenssen-Erz, T. (2007). Cattle herding in the southern Libyan Desert. In Bubenzer, O., Bolten, A., Darius F. (eds.) Atlas of Cultural and Environmental Change in Arid Africa. pp. 46-49. Köln, Germany: Heinrich-Barth Institut. Kienast, F., Degenhardt, B., Weilenmann, B., Wäger, Y., Buchecker, M. (2012). GIS-assisted mapping of landscape suitability for nearby recreation. Landscape and Urban Planning, 105(4): 385–399. doi: https://doi.org/10.1016/j.landurbplan.2012.01.015. Kirchhoff, T. (2019). Abandoning the concept of cultural ecosystem services, or against natural- scientific imperialism. BioScience: biz007. doi: https://doi.org/10.1093/biosci/biz007. Komossa, F., van der Zanden, E.H., Schulp, C.J.E., Verburg, P.H. (2018). Mapping landscape potential for outdoor recreation using different archetypical recreation user groups in the European Union. Ecological Indicators, 85: 105-116. doi: 10.1016/j.ecolind.2017.10.015. Kubiszewski, I., Costanza, R., Anderson, S., Sutton, P. (2017). The future value of ecosystem services: Global scenarios and national implications. Ecosystem Services, 26: 289–301. doi: 10.1016/j.ecoser.2017.05.004. Lanouar, C., Goaied, M. (2019). Tourism, terrorism and political violence in Tunisia: Evidence from Markov-switching models. Tourism Management, 70: 404–418. doi: 10.1016/j.tourman.2018.09.002. Lemelin, H., Dawson, J., Stewart, E. J. (2012). Last-chance tourism: Adapting tourism opportunities in a changing world. London, UK: Routledge. Le Quellec, J.-L. (2009). Les images rupestres du Jebel el-’Uweynât. Archéo-Nil, 19: 13-26. FCUP 169 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Le Quellec, J.-L. (2013). Aréologie, phénétique et art rupestre: l’exemple des théranthropes du Sahara central. Les Cahiers de l’AARS, 16: 155-76. Levin, N., Ali, S., Crandall, D., Kark, S. (2019). World Heritage in danger: Big data and remote sensing can help protect sites in conflict zones. Global Environmental Change, 55: 97–104. doi: 10.1016/j.gloenvcha.2019.02.001. Lluch, P., Philip, P. (2003). Six stations à gravures du N.E. de l'Adrar (Dhar Chinguetti, Mauritanie). Cahiers de l’AARS, 8: 87-96. Lu, N., Wang, M., Ning, B., Yu, D., Fu, B. (2018). Research advances in ecosystem services in drylands under global environmental changes. Current Opinion in Environmental Sustainability, 33: 92–98. doi: 10.1016/j.cosust.2018.05.004. Maes, J. Egoh, B., Willemen, L. Liquete, C. Vihervaara, P. Schägner, J. P. Grizzetti, B. Drakou, E.G., La Notte, A., Zulian, G., Bouraoui, F., Paracchinia, M.L., Braat, L., Bidoglio, G. (2012). Mapping ecosystem services for policy support and decision making in the European Union. Ecosystem Services, 1(1): 31-39. doi: https://doi.org/10.1016/j.ecoser.2012.06.004. Martinez-Harms, M.J., Bryan, B.A., Balvanera, P., Law, E.A., Rhodes, J.R., Possingham, H.P., Wilson, K.A. (2015). Making decisions for managing ecosystem services. Biological Conservation, 184: 229–238. doi: 10.1016/j.biocon.2015.01.024. Martínez-Harms, M.J., Balvanera, P. (2012). Methods for mapping ecosystem service supply: a review. International Journal of Biodiversity Science, Ecosystem Services & Management, 8(1- 2): 17–25. doi: 10.1080/21513732.2012.663792. Milcu, A.I., Hanspach, J., Abson, D., Fischer, J. (2013). Cultural ecosystem services: A literature review and prospects for future research. Ecology and Society, 18(3): 44. Accessed on 06 February 2019 at http://dx.doi.org/10.5751/ES-05790-180344. Millennium Ecosystem Assessment (MEA) (2005). Ecosystems and Human Well-being: Synthesis. Island Press: Washington, DC. Moilanen, A., Pouzols, F.M., Meller, L., Veach, V., Arponen, A., Leppänen, J., Kujala, H. (2014). Zonation spatial conservation planning framework and software v. 4.0, User Manual. (University of Helsinki, 2014). http://cbig.it.helsinki.fi/files/zonation/zonation_manual_v4_0.pdf. Accessed on 05 June 2019. Nahuelhual, L., Carmona, A., Lozada, P., Jaramillo, A., Aguayo, M. (2013). Mapping recreation and ecotourism as a cultural ecosystem service: An application at the local level in Southern Chile. Applied Geography, 40: 71–82. doi: https://doi.org/10.1016/j.apgeog.2012.12.004.

170 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Nahuelhual, L., Vergara, X., Kusch, A., Campos, G., Droguett, D. (2017). Mapping ecosystem services for marine spatial planning: Recreation opportunities in Sub-Antarctic Chile. Marine Policy, 81: 211–218. doi: https://doi.org/10.1016/j.marpol.2017.03.038. Naidoo, R., Balmford, A., Costanza, R., Fisher, B., Green, R.E., Lehner, B., Malcolm, T.R., Ricketts, T.H. (2008). Global mapping of ecosystem services and conservation priorities. PNAS, 105(28): 9495-9500. doi: https://doi.org/10.1073/pnas.0707823105. NGA (2016). National Geospatial-Intelligence Agency GEOnet Names Server (GNS). Accessed on 15 November of 2018 at http://geonames.nga.mil/gns/html. NIMA (1997). Vector Map (VMap) Level 0. National Imagery and Mapping Agency's (NIMA). Accessed on 06 September 2018 at http://earth-info.nga.mil/publications/vmap0.html. Noguera, A.M., Zboray, A. (2011). Containers, bags, and other manmade objects in the Pastoral paintings of the Jebel el-‘Uweinat: A review. Cahiers de l’AARS, 15: 275-294. OECD-SWAC (2014). An Atlas of the Sahara-Sahel. Geography, Economics and Security. OECD Publishing. Accessed at http://dx.doi.org/10.1787/9789264222359-en. O’Farrell, P.J., Reyers, B., Le Maitre, D.C., Milton, S.J., Egoh, B., Maherry, A., Colvin, C., Atkinson, D., De Lange, W., Blignaut, J.N., Cowling, R.M. (2010). Multi-functional landscapes in semi arid environments: implications for biodiversity and ecosystem services. Landscape Ecology, 25(8): 1231–1246. doi: 10.1007/s10980-010-9495-9. Oteros-Rozas, E., Martín-López, B., Fagerholm, N., Bieling, C., Plieninger, T. (2018). Using social media photos to explore the relation between cultural ecosystem services and landscape features across five European sites. Ecological Indicators, 94(2): 74-86. https://doi.org/10.1016/j.ecolind.2017.02.009. Pandeya, B., Buytaert, W., Zulkafli, Z., Karpouzoglou, T., Mao, F., Hannah, D.M. (2016). A comparative analysis of ecosystem services valuation approaches for application at the local scale and in data scarce regions. Ecosystem Services, 22: 250–259. doi: 10.1016/j.ecoser.2016.10.015. Paracchini, M. L., Zulian, G., Kopperoinen, L., Maes, J., Schägner, J. P., Termansen, M., Zandersen, M., Perez-Sobad, M., Scholefield, P.A., Bidoglio, G. (2014). Mapping cultural ecosystem services: A framework to assess the potential for outdoor recreation across the EU. Ecological Indicators, 45: 371–385. doi: https://doi.org/10.1016/j.ecolind.2014.04.018.2 Peña, L., Casado-Arzuaga, I., Onaindia, M. (2015). Mapping recreation supply and demand using an ecological and a social evaluation approach. Ecosystem Services, 13: 108–118. doi: https://doi.org/10.1016/j.ecoser.2014.12.008. FCUP 171 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Piggott-McKellar, A.E. and McNamara, K.E. (2016). Last chance tourism and the Great Barrier Reef. Journal of Sustainable Tourism, 25(3): 397-415. doi: https://doi.org/10.1080/09669582.2016.1213849. Plieninger, T., Dijks, S., Oteros-Rozas, E., Bieling, C. (2013). Assessing, mapping, and quantifying cultural ecosystem services at community level. Land Use Policy, 33: 118–129. doi: 10.1016/j.landusepol.2012.12.013. Powers, R.P., Jetz, W. (2019). Global habitat loss and extinction risk of terrestrial vertebrates under future land-use-change scenarios. Nature Climate Change, 9: 323-329. doi: https://doi.org/10.1038/s41558-019-0406-z. PSSC (2018). Earth Impact Database. Planetary and Space Science Centre, Department of Geology, University of New Brunswick. Accessed on 06 September 2018 at http://www.passc.net/EarthImpactDatabase/. R Core Team (2017). R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org/. Raleigh, C., Linke, A., Hegre, H., Karlsen, J. (2010). Introducing ACLED-Armed conflict location and event data. Journal of Peace Research, 47: 651-660. doi: https://doi.org/10.1177/0022343310378914. Rekacewicz, P. (2012). Sahara-Sahel: movements and routes. Le Monde Diplomatique March 2012. Accessed at http://mondediplo.com/maps/saharasahel. Rieb, J.T., Chaplin-Kramer, R., Daily, G.C., Armsworth, P.R., Böhning-Gaese, K., Bonn, A., Cumming, G.S., Eigenbrod, F., Grimm, V., Jackson, B.M., Marques, A., Pattanayak, S.K., Pereira, H.M., Peterson, G.D., Ricketts, T.H., Robinson, B.E., Schröter, M., Schulte, L.A., Seppelt, R., Turner, M.G., Bennett, E.M. (2017). When, where, and how nature matters for ecosystem services: Challenges for the next generation of ecosystem service models. BioScience, 67(9): 820-833. doi: https://doi.org/10.1093/biosci/bix075. Riemer, H. (2009). Prehistoric rock art research in the western desert of Egypt. Archéo-Nil, 19: 31–46. Rosa, I.M.D., Pereira, H.M., Ferrier, S., Alkemade, R., Acosta, L.A., Akcakaya, H.R., den Belder, E., Fazel, H.M., Fujimori, S., Harfoot, M., Harhash, K.A., Harrison, P.A., Hauck, J., Hendriks, R.J.J., Hernández, G., Jetz, W., Karlsson-Vinkhuyzen, S.I., Kim, H., King, N., Kok, M.T.J., Kolomytsev, G.O., Lazarova, T., Leadley, P., Lundquist, C.J., Márquez, J.G., Meyer, C., Navarro, L.M., Nesshöver, C., Ngo, H.T., Ninan, K.N., Palomo, M.G., Pereira, L.M., Peterson, G.D., Pichs, R., Popp, A., Purvis, A., Ravera, F., Rondinini, C., Sathyapalan, J., Schipper, A.M., Seppelt, R., Settele, J., Sitas, N., van Vuuren, D. (2017). Multiscale scenarios for nature

172 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

futures. Nature Ecology & Evolution, 1: 1416-1419. doi: https://doi.org/10.1038/s41559-017- 0273-9. Saarinen, J., 2016. Political ecologies and economies of tourism development in Kaokoland, North-West Namibia. In: Mostafanezhad, M., Carr, A., Norum, R. (Eds), Political Ecology of Tourism: Communities, Power and the Environment. pp. 213-230. Routledge: London. Saarinen, J. (2018). What are wilderness areas for? Tourism and political ecologies of wilderness uses and management in the Anthropocene. Journal of Sustainable Tourism. doi: https://doi.org/10.1080/09669582.2018.1456543. Saarinen, J., Moswete, N., Monare, M.J. (2014) Cultural tourism: new opportunities for diversifying the tourism industry in Botswana. Bulletin of Geography: Socio-Economic Series, 26(26): 7-18. doi: https://doi.org/10.2478/bog-2014-0041. Safriel, U., Adeel, Z. (2005). Dryland systems: Ecosystems and Human Well-being. In Hassan, R. Scholes, Neville, A. (eds.) Current State and Trends, Vol. 1. pp. 623-662. Island Press: Washington, District of Columbia, USA. Santarém, F., Campos, J., Pereira, P., Hamidou, D., Saarinen, J., Brito, J.C. (2018). Using multivariate statistics to assess ecotourism potential of water-bodies: A case-study in Mauritania. Tourism Management, 67: 34-46. doi: https://doi.org/10.1016/j.tourman.2018.01.001. Santarém, F., Pereira, P., Saarinen, J., Brito, J.C. (2019a). New method to identify and map flagship fleets for promoting conservation and ecotourism. Biological Conservation, 229: 113- 124. doi: https://doi.org/10.1016/j.biocon.2018.10.017. Santarém, F., Saarinen, J., Brito, J.C. (2019). Desert Conservation and Management: Ecotourism. In: Goldstein MI, DellaSala DA (Eds.), Encyclopedia of the World's Biomes, vol. 2. Elsevier, pp. 259-273. doi: https://doi.org/10.1016/B978-0-12-409548-9.11827-5. Schulp, C.J.E., Burkhard, B., Maes, J., Van Vliet, J., Verburg, P.H. (2014). Uncertainties in ecosystem service maps: A comparison on the European scale. PLoS ONE, 9(10): e109643. doi: 10.1371/journal.pone.0109643. Scholte, S.S.K., Daams, M., Farjon, H., Sijtsma, F.J., van Teeffelen, A.J.A., Verburg, P.H. (2018). Mapping recreation as an ecosystem service: Considering scale, interregional differences and the influence of physical attributes. Landscape and Urban Planning, 175: 149–160. doi: 10.1016/j.landurbplan.2018.03.011. Seppelt, R., Dormann, C. F., Eppink, F. V., Lautenbach, S., Schmidt, S. (2011). A quantitative review of ecosystem service studies: approaches, shortcomings and the road ahead. Journal of Applied Ecology, 48(3): 630–636. doi: 10.1111/j.1365-2664.2010.01952.x. FCUP 173 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Small, N., Munday, M., Durance, I., (2017). The challenge of valuing ecosystem services that have no material benefits. Global Environmental Change, 44: 57-67. doi: http:/dx.doi.org/10.1016/j.gloenvcha.2017.03.005. START (2015). Global Terrorism Database. National Consortium for the Study of Terrorism and Responses to Terrorism. Accessed on 12 December 2018 at http://www.start.umd.edu/gtd. Stålhammar, S., Pedersen, E. (2017). Recreational cultural ecosystem services: How do people describe the value? Ecosystem Servises, 26: 1-9. doi: https://doi.org/10.1016/j.ecoser.2017.05.010 TEEB (2010). The Economics of Ecosystems and Biodiversity: Mainstreaming the Economics of Nature: A Synthesis of the approach, conclusions and recommendations of TEEB. Accessed on 04 February of 2019 at http://www.teebweb.org/. United Nations Educational, Scientific and Cultural Organization (UNESCO) (2018). World Heritage List. UNESCO World Heritage Centre. Accessed on 15 November of 2018 at https://whc.unesco.org/en/list/. United Nations Environmental Programme (UNEP). (2006a). Global deserts outlook (ed. E. Ezcurra). Accessed on 14 February 2019 at http://wedocs.unep.org/handle/20.500.11822/9581. United Nations Environmental Programme (UNEP). (2006b). Tourism and deserts: A practical guide to managing the social and environmental impacts in the desert recreation sector (eds. de Assis, H. R., Manca M.). Accessed on 14 February 2019 at http://wedocs.unep.org/handle/20.500.11822/2229. United Nations Development Programme (UNDP) (2018). Human Development Report 2018. Accessed on 05 February of 2019 at http://www.hdr.undp.org/. United Nations World Tourism Organization (UNWTO). (2018a). Promotion of sustainable tourism, including ecotourism, for poverty eradication and environment protection. Accessed on 04 February of 2019 at http://sdt.unwto.org/unga-sustainable-tourism-resolutions. United Nations World Tourism Organization (UNWTO). (2018b). Tourism Highlights: 2018 Edition. Madrid: UNWTO. Accessed on 04 February of 2019 at http://www2.unwto.org/. doi: https://doi.org/10.18111/9789284419876. United Nations World Tourism Organization (UNWTO)., United Nations Development Programme (UNDP). (2017). Tourism and the Sustainable Development Goals – Journey to 2030. Accessed on 04 February of 2019 at http://www2.unwto.org/publication/tourism-and- sustainable-development-goals-journey-2030.

174 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

USGS. (2018). Shuttle Radar Topography Mission (SRTM). United States Geological Survey. Accessed on 06 September 2018 at https://lta.cr.usgs.gov/SRTM. Vale, C.G., Pimm, S.L., Brito, J.C. (2015). Overlooked mountain rock pools in deserts are critical local hotspots of biodiversity. PLoS ONE, 10(2): e0118367. doi: https://doi.org/10.1371/journal.pone.0118367. van Berkel, D.B., Verburg, P. H. (2014). Spatial quantification and valuation of cultural ecosystem services in an agricultural landscape. Ecological Indicators, 37: 163–174. doi: 10.1016/j.ecolind.2012.06.025. van Zanten, B.T., Van Berkel, D.B., Meentemeyer, R.K., Smith, J.W., Tieskens, K.F., Verburg, P.H. (2016). Continental-scale quantification of landscape values using social media data. PNAS, 113(46): 12974. doi: www.pnas.org/cgi/doi/10.1073/pnas.1614158113. Vaz, A.S., Gonçalves, J.F., Pereira, P., Santarém, F., Vicente, J.R., Honrado, J.P. (2019). Earth observation and social media: Evaluating the spatiotemporal contribution of non-native trees to cultural ecosystem services. Remote Sensing of Environment, 230: 111193. doi: 10.1016/j.rse.2019.05.012. Wall, J., Wittemyer, G., Klinkenberg, B., LeMay, V., Douglas-Hamilton, I. (2013) Characterizing properties and drivers of long distance movements by elephants (Loxodonta africana) in the Gourma, Mali. Biological Conservation, 157: 60-68. doi: https://doi.org/10.1016/j.biocon.2012.07.019. Waldron, A., Mooers, A.O., Miller, D.C., Nibbelink, N., Redding, D., Tyler, S.K., Timmons Roberts, J., Gittleman, J.L. (2013). Targeting global conservation funding to limit immediate biodiversity declines. PNAS, 110: 12144-12148. doi: https://doi.org/10.5061/dryad.p69t1. Watson, J.E.M., Venter, O., Lee, J., Jones, K.R., Robinson, J.G., Possingham, H.P., Allan, J.R. (2018). Protect the last of the wild. Nature, 563: 27-30. doi: 10.1038/d41586-018-07183-6. Wei, H., Liu, H., Xu, Z., Ren, J., Lu, N., Fan, W., Zhang, P., Dong, X. (2018). Linking ecosystem services supply, social demand and human well-being in a typical mountain–oasis–desert area, Xinjiang, China. Ecosystem Services, 31: 44–57. doi: 10.1016/j.ecoser.2018.03.012. Weiss, C. (2016). Al Qaeda has launched more than 100 attacks in West Africa in 2016. The Long War Journal. Accessed at http://www.longwarjournal.org/archives/2016/06/over-100-al-qaeda- attacks-in-west-africa-since-beginning-of-the-year.php. Weyland, F., Laterra, P. (2014). Recreation potential assessment at large spatial scales: A method based in the ecosystem services approach and landscape metrics. Ecological Indicators, 39: 34-43. doi: https://doi.org/10.1016/j.ecolind.2013.11.023. FCUP 175 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Wolff, S., Schulp, C.J.E., Verburg, P.H. (2015). Mapping ecosystem services demand: A review of current research and future perspectives. Ecological Indicators, 55: 159-171. doi: https://doi.org/10.1016/j.ecolind.2015.03.016. Wood, S.L., Jones, S.K., Johnson, J.A., Brauman, K.A., Chaplin-Kramer, R., Fremier, A., Girvetz, E., Gordon, L.J., Kappel, C.V., Mandle, L. (2018). Distilling the role of ecosystem services in the Sustainable Development Goals. Ecosystem Services, 29: 70-82. doi: https://doi.org/10.1016/j.ecoser.2017.10.010. Zeppel, H.D. (ed.) (2006). Indigenous Ecotourism: Sustainable Development and Management. 308pp. Wallingford, UK: CAB International.

176 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Chapter 5

Ecosystem services management

Article V Santarém, F., Saarinen, J., Brito, J.C. (2021). “Assessment and prioritization of cultural ecosystem services in the Sahara-Sahelian region”. Science of the Total Environment, 777: 146053 Article VI Santarém, F., Gonçalves, D., Saarinen, J., Brito, J.C. (under review). “Vulnerability of desert tourism to climate change: lessons from the Sahara- Sahel”. Journal of Sustainable Tourism

FCUP 177 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Article V: Assessment and prioritization of cultural ecosystem services in the Sahara- Sahelian region

Frederico Santarém1,2,3, Jarkko Saarinen3,4, José Carlos Brito1,2

1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 - Geography Research Unit, University of Oulu, Finland 4 - School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

178 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Abstract

Desert environments remain largely neglected by the society and their potential to provide benefits to people remain understudied. Hotspots of cultural ecosystem services have been identified in some deserts; yet, knowing which countries need to strengthen efforts to satisfy people’s demand for those services is timely needed. Here, we show the performance of countries within the Earth’s largest warm region – the Sahara-Sahel – in managing cultural ecosystem services. Using the most-advanced decision-support tools and updated databases on biodiversity features and constrains to ecosystem services and on socioeconomic indicators, we identified national priorities for cultural services management. We also identified countries that are missing opportunities for local sustainable development. About 34% of Sahara-Sahel is prioritized for cultural ecosystem services, particularly in the main mountains and waterbodies of the region and along the Western and Eastern coastal limits. Algeria, Egypt, Libya, Morocco, Senegal, and Tunisia are performing better in managing their cultural services given the availability of such services in their territories. Burkina Faso, Cameroon, Chad, Egypt, Libya, Mali, Niger, Nigeria, Sudan, and South Sudan need to urgently improve their ease of mobility, governance, safety, socioeconomic and health systems to foster ecosystem services demand. Cameroon, Eritrea, and Senegal are receiving far less tourists than what their ecosystems can handle and need to improve their local conditions for better marketing international tourists able to economically contribute to sustainable development through ecotourism programs. The approach developed here serves as a framework for conserving the last world wild ecosystems and is replicable to other contexts where regional planning for ecosystem management is compulsory.

Keywords Sahara-Sahelian Region; Cultural Ecosystem Services; Maximum Likelihood Classification; sustainable development; Principal Component Analysis; socio-economic analyses FCUP 179 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Introduction

Global ecosystems are imperiled by unsustainable human development and conflict (Cardinale et al., 2012; Díaz et al., 2019; Levin et al., 2019; Maron et al., 2019; Newbold et al., 2015; Tilman et al., 2017). The largest warm region of the world (≈11,200,000 km2), which includes the African ecoregions of the Sahara Desert and its contiguous Sahel arid area (Fig. D1; Dinerstein et al., 2017), is no exception and several human factors are dismantling the region’s ecosystems. Armed conflicts threaten biodiversity and local people (Brito et al., 2018; Daskin & Pringle, 2018; Dioko & Harrill, 2019; Walther, 2017), and the unsustainable use of natural resources severely depletes natural ecosystems (Brito et al., 2014, 2018; OECD/SWAC, 2014; Santarém et al., 2020b). Furthermore, while many countries within Sahara-Sahel will see a large proportion of their biodiversity to be up-listed in conservation status (Durant et al., 2014; Powers & Jetz, 2019) they are among the least studied (Brito et al., 2014; Durant et al., 2012, 2014) and among the most underfunded for conservation worldwide (Waldron et al., 2013). The region is among the last natural places on earth (Di Marco et al., 2019; Watson et al., 2018) partially managed by indigenous people (Garnett et al., 2018). This offers chances to tourists seeking ‘last- chance-to-see’ wild places (Saarinen, 2019) to benefit from the many cultural ecosystem services (CES) that the region supplies (Santarém et al., 2020b) (CES refer to non-material benefits that people obtain from nature through experiences, recreation and tourism, for example; Díaz et al., 2019). Most Sahara-Sahel countries remain among the poorest (UNDP, 2019) and least visited worldwide (UNWTO, 2019), fostering a continuing regional ecosystem crisis (Brito et al., 2018), and yet the region remains neglected by the global society (Durant et al., 2012). To counteract the degradation of ecosystems and to improve human wellbeing worldwide, CES have received an increasing importance in research and policy making (Díaz et al., 2018; Kosanic & Petzold, 2020; Wood et al., 2018). Among the many CES that desert environments can provide (see Safriel et al., 2005 and Teff‐Seker & Orenstein, 2019), ecotourism and related recreational activities have been proposed as key to combat poverty and stimulate environmental protection (Santarém et al., 2020a; Santarém & Paiva, 2015; UNEP, 2006b; Winkler & Brooks, 2020). To achieve sustainability in deserts (or elsewhere), it is crucial that decision-makers contemplate the spatial characteristics of CES supply and management (Burkhard et al., 2012), but analyzing CES is challenged by their intangibility (Cheng et al., 2019; Ogada et al., 2012; Satz et al., 2013; Small et al., 2017; Teoh et al., 2019). Even considering this challenge, there is a growing body of research concerned with mapping CES worldwide (Bachi et al., 2020; Casado- Arzuaga et al., 2014; He et al., 2019; Nahuelhual et al., 2013; Peña et al., 2015; Scholte et al.,

180 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

2018), with several methods being developed to date, such as ESTIMAP (Zulian et al., 2014), ARIES (Villa et al., 2014), or InVEST (Sharp et al., 2014). Several methodologies for ecosystem services assessments and their limitations have been revised in the ecosystem services literature (Brown & Fagerholm, 2015, Crossman et al., 2013, Kosanic & Petzold, 2020, Martínez-Harms & Balvanera 2012, Milcu et al., 2013 and Wolff et al., 2015). In this regard, desert CES have been recently assessed and mapped (Santarém et al., 2020b; Taylor et al., 2017; Teff‐Seker & Orenstein, 2019; Winkler & Brooks, 2020), but very few studies are concerned with the management of ecosystem services in North African deserts (Hosni, 2000; Santarém et al., 2020a, 2020b; UNESCO, 2003b, 2003a; Vale et al., 2015). Previous studies identified many CES hotspots within Sahara-Sahel (Santarém et al., 2020b) but neglected the broad-scale socio-economic conditions that underly ecosystem management in the region. These CES-rich regions remain poorly protected (Brito et al., 2016) and are suggested to face increased environmental degradation from growing human population and wealth (OECD/SWAC, 2014). Given these facts, knowing which Sahara-Sahel countries are doing better than their peers in managing CES is timely needed. Solving all these issues is of paramount importance for regional policymakers to manage the various elements of human wellbeing in one of the Earth’s most threatened regions and to provide the elements for sustainability (Blicharska et al., 2019; Wood et al., 2018). Globally, developed countries are associated with higher visitation levels (UNWTO, 2019). In a relative sense, countries within Sahara-Sahel show similar patterns, with the most economically developed ones being more visited than the less developed states (Fig. D2). Countries with biodiversity hotspots tend to have higher annual growth of tourism investments than places without ecosystem-based attractions (Blicharska et al., 2019) and, in Africa, especially in eastern and southern regions, ecotourism is commonly promoted as a tool for both conservation and socioeconomic development (World Tourism Organization, 2014). However, we lack information about which Sahara-Sahel countries are potentially missing opportunities to attract ecotourists interested in supporting nature conservation and improving local economies in deserts. Knowing this is needed to improve countries commitment to CES management, as local governments can be further pressed to maintain pristine ecosystems so international tourists are attracted and contribute financially to the sustainable development of the region. Here, we address all these knowledge gaps and provide the first estimate of countries performance in managing CES in the largest warm region in the world, the Sahara-Sahel. Furthermore, we identify which countries are missing opportunities to use and promote desert CES for sustainable development by using ecotourism, given the availability of areas supplying FCUP 181 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

CES that are priorities for conservation within their territories. Our analyses are indicative at the regional and national scales and form a usable framework for conserving the last wild ecosystems in the world (Di Marco et al., 2019), including the Sahara-Sahel region. We addressed four questions that are fundamental for improving our understanding of the importance of desert CES and for effectively implementing strategies to both promote and preserve desert ecosystems under threat: (1) among the lands supplying the highest levels of CES, which ones are priority for conservation and management?; (2) which socioeconomic variables are constraining national conditions for better CES management?; (3) given the prioritized lands of CES availability and the socioeconomic conditions within each Sahara-Sahel country, what is their performance in managing CES?; and (4) which countries display substantial availability of CES and yet are missing opportunities for socioeconomic development based on ecotourism? There are some data and methodological limitations in answering fully to all these research questions, which are discussed in detail later.

Methods

We conducted an interdisciplinary analysis with several steps, from spatial data collection to spatial prioritization, and from socioeconomic data assortment to evaluation of the performance of countries in managing CES in the Sahara-Sahel (Fig. D3). We constructed environmental performance indicators specific to Sahara-Sahel nations by making use of the most updated and high-resolution data available and advanced decision-support tools developed to date (Moilanen et al., 2014). These tools allowed us to perform a prioritization ranking of the Sahara-Sahel landscapes for CES management among different countries. We gathered revised spatially- explicit and fine-scale data related to CES in deserts (Table 4.1; Santarém et al., 2020a), followed by related prioritization results with the most updated socioeconomic indicators (Table 4.3) to evaluate the performance of Sahara-Sahel countries in managing their CES. We also identified countries missing opportunities for socioeconomic development through operational ecotourism by relating inbound tourist data with the number of prioritized landscape units for CES conservation.

182 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Spatial Component

Spatial data processing To perform a spatial prioritization of the landscape for CES management, one needs to first gather relevant attraction and constraint features to conservation (Moilanen et al., 2014). We collected a set of 148 spatially-explicit data layers representative of 24 variables influencing CES in deserts (Santarém et al., 2020b, 2020a; UNEP, 2006b, 2006a) from multiple sources (Table 4.1). In all layers, input data were converted to the geographic coordinate system WGS84 and rasterized to a 0.5º spatial resolution (n = 4120 pixels of 50km2), using the geographic information system ArcGIS 10.1 (ESRI, 2012). The layers consisted of positive features: species (2 variables: large- and small-flagships), other biodiversity (3 variables: forest reserves, protected areas and UNESCO World Heritage Sites), landscape (8 variables: major landscape features, gorges and mountain passes, peculiar rock formations, caves, major wetlands, mountain rock pools, desert ecosystem intactness, and areas of extreme remoteness), and cultural (7 variables: Sahara-Sahel ethnographic groups, oases, monuments, caravan villages, fortifications from the colonial period, sites of historical occupation and rock art), and of cost features: conflict (2 variables: landmines and attacks and violence against civilians) and exploitation of natural resources (2 variables: oil, gas, mining facilities, and high Human Footprint) on Sahara and Sahel ecoregions (see Santarém et al., 2020a for methodological details on spatial data acquisition and processing). A detailed explanation on the spatial data acquisition and processing is available in Supplementary Material (Text D1).

Table 4.1. Categories and subcategories of the variables used in the prioritization rankings, the number of data layers within each variable, the approximate feature-specific numerical priority weight of each feature, and data sources. See references for data details.

Category Sub-category Variable Data layers Weight value Data source

Features Species Large flagships 33 0.015 Santarém et al. (2019)

Small flagships 69 0.007 Santarém et al. (2019)

Other Forest reserves 1 0.333 Santarém et al. Biodiversity (2020a)

Protected Areas 1 0.333 Santarém et al. (2020a) FCUP 183 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

UNESCO World Heritage Sites 1 0.333 Santarém et al. (2020a)

Landscape Major landscape features 12 0.010 Santarém et al. (2020a)

Gorges and mountain passes 1 0.125 Santarém et al. (2020a)

Peculiar rock formations 1 0.125 Santarém et al. (2020a)

Caves 1 0.125 Santarém et al. (2020a)

Major wetlands 4 0.031 Santarém et al. (2020a)

Mountain rock pools 1 0.125 Santarém et al. (2020a)

Human Footprint (Low) 1 0.125 Venter et al. (2018, 2016a)

Extreme remoteness 1 0.125 Venter et al. (2018, 2016a)

Cultural Major ethnographic groups 11 0.013 Santarém et al. (2020a)

Oases 1 0.143 Santarém et al. (2020a)

Monuments 1 0.143 Santarém et al. (2020a)

Caravan villages 1 0.143 Santarém et al. (2020a)

Fortifications from colonial period 1 0.143 Santarém et al. (2020a)

Ruins, tombs, sites of historical land 1 0.143 Santarém et al. occupation (e.g., Roman, Ancient (2020a) Egypt)

Rock art 1 0.143 Santarém et al. (2020a)

Costs Conflict Landmines 1 -0.321 Santarém et al. (2020a)

184 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Attacks/battles and violence against 1 -0.642 Santarém et al. civilians (2011-2016) (2020a)

Exploitation of Oil, gas, mining facilities 1 -0.032 Santarém et al. natural (2020a) resources

Human Footprint (High) 1 -0.006 Venter et al. (2018, 2016)

Lock out Conflict Berm in territory 1 - United Nations (2019)

Spatial prioritization process We used the most recent publicly available decision-support tool, Zonation v4.0 (Lehtomäki & Moilanen, 2013; Moilanen et al., 2014) to produce ranking maps for CES in Sahara- Sahel (n = 4120 pixels of 0.5º spatial resolution). This software is capable of processing problems of four or more orders of magnitude bigger than previously possible (Kremen et al., 2008; Lehtomäki et al., 2009; Lehtomäki & Moilanen, 2013; Pouzols et al., 2014). Zonation uses a maximum-coverage approach, aiming to maximize the conservation benefits for a fixed cost (Moilanen et al., 2011). It prioritizes maps by ranking landscape elements (as pixels) iteratively from the lowest to the highest priority for conservation (Lehtomäki & Moilanen, 2013). The removal order thus reflects the rank order of importance of planning units to the systematic planning issue, with the most important units remining until last (Moilanen, 2007; Moilanen et al., 2005, 2014). Zonation produces a balanced ranking, meaning that for any given rank level areas are complementary, and jointly achieve a well-balanced representation across all biodiversity features (Moilanen et al., 2014; Pouzols et al., 2014), a key objective in systematic planning (Kukkala & Moilanen, 2013). We used the additive benefit function (ABF) analysis variant of Zonation, which can be interpreted as the "minimization of aggregate extinction rates via feature-specific species-area curves" (Moilanen et al., 2014). This means that ABF favors grid cells containing large numbers of localized features, summing values across features in each cell (Arponen et al., 2005; Moilanen, 2007). The ABF variant has been extensively used in large spatial analyses (Pouzols et al., 2014) as it produces high return-on-investment solutions (Laitila & Moilanen, 2012; Pouzols et al., 2014), by retaining higher fractions of features’ distributions without requiring arbitrary targets (Di Minin & Moilanen, 2012). Additionally, the ABF allows to use cost layers without compromising efficiency and interpretation of prioritization results (Moilanen et al., 2014).

FCUP 185 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Variables weighting During the ranking process, Zonation uses feature-specific numerical priority weights that influence the relative balance that emerges between features in the final solution (Arponen et al., 2005; Leathwick et al., 2008; Lehtomäki & Moilanen, 2013). To avoid unequal sub-category weights based on the different number of variables within each sub-category (e.g. an aggregate weight of 271 for biodiversity and 37 for landscape), which would potentially lead to the sub- category with the largest number of variables and data layers having the greatest influence on the analyses outcomes, we set the same combined weight to each sub-category and rescaled the weights of each individual data layer to sum up to the aggregate sub-category group weight (see categories and weight values in Table 4.1). This process ensured a flexible and balanced weighting, while guaranteeing that the software would treat all features and costs equally during the ranking process (Di Minin et al., 2017; Lehtomäki & Moilanen, 2013). The success of an informed prioritization exercise depends on the thoroughly consideration of constraints to CES as representative costs to conservation actions (Arponen et al., 2010; Naidoo et al., 2006; Whitehead et al., 2014). Applying negative weights to cost features ensures that areas hosting those features are removed early in the prioritization, while positive features are retained to the top ranks of the landscape (Moilanen et al., 2014). Thus, negative weights were given to the cost features constraining CES (Di Minin et al., 2017; Kujala et al., 2018; Moilanen et al., 2011, 2014; Whitehead et al., 2014). We performed a cost sensitivity analysis, by testing three cost weightings (see Fig. D4): costs were weighted one time (1x), one- hundred times (100x), and one-thousand times (1000x). Because costs tend to dominate the solution when higher multiplying factors are applied (Moilanen et al., 2014), we conducted further analyses using the simplest approach (1x cost). By weighting the aggregated set of costs equally to the aggregated weights of features (i.e., making sure costs in aggregate would sum one), we ensured a balanced solution between features and costs (Di Minin et al., 2017). Still, the integration of costs requires careful consideration, as other factors may influence the effectiveness of spatial prioritizations, such as dynamic social and economic factors (Arponen et al., 2010; Pouzols et al., 2014; Waldron et al., 2013), especially in African countries subjected to weak governance (Bradshaw & Di Minin, 2019).

Ensemble prioritization framework Despite ecosystem services are supplied along the continuum of landscape and do not recognize artificial human-imposed borders, priorities for ecosystem management can vary between regions (Moilanen & Arponen, 2011). In systematic planning exercises, global

186 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

prioritizations retrieve different, and sometimes conflicting, outputs in comparison to national ones (Kukkala et al., 2016; Moilanen et al., 2013; Moilanen & Arponen, 2011; Pouzols et al., 2014). Strong evidence suggests an efficiency loss on prioritizations from global to national or local solutions (Kukkala et al., 2016; Moilanen & Arponen, 2011; Pouzols et al., 2014). Yet, alternative prioritization scenarios present distinct methodological details that need to be thoroughly considered. For example, some variant analyses may highlight country-specific priorities (using the strong administrative priorities analysis option), others may emphasize global solutions (see Moilanen et al. (2014) and Moilanen and Arponen (2011) for details), and others may produce “edge effects” when administrative boundaries artificially cut the distribution of features, affecting the outcome of prioritization solutions (Moilanen et al., 2013). Combining continental and national analyses has the potential to increase decision-making efficiency, by enabling country-level decision-making to occur in the context of international priorities (Moilanen et al., 2013; Pouzols et al., 2014). Thus, to account for uncertainties in the prioritization ranking, we conducted 10 prioritizations with varying settings (Whitehead et al., 2014) (see the different scenarios in Fig. D4 and Table 4.2). When needed, the area of countries was used as input region-specific weights (Moilanen et al., 2014; Moilanen & Arponen, 2011) (Table 4.2). Country areas were calculated with the function Calculate geometry in ArcMap 10.1 (WGS 1984 World Mercator) (ESRI, 2012), and were normalized to aggregately sum to one (Moilanen et al., 2013). Administrative country boundaries were based in the Global Administrative Areas spatial database (GADM, 2018), adjusted for their limits within Sahara-Sahel ecoregions (Dinerstein et al., 2017). The 10 prioritization scenarios were scaled onto a 3 classes score scale according to the prioritization ranking for CES availability (Fig. D4). We ensembled the 10 prioritization outputs into an averaged priority rank (Meller et al., 2014) and scaled it onto a 1-3 score scale, which was used in the following analyses (Fig. 4.1).

Table 4.2. List of the 10 scenarios used in the ranking process. Considerations of costs (weighted at 1x, 100x and 1000x times; see Fig. D4), administrative units, administrative priorities, global weights, and normalization of country area are specified. See Moilanen et al. (2014) for complete methodological details of each scenario.

Administrative Administrative Global Area of country Scenario Cost Units priority weight normalized a NO NO – – NO b YES NO – – NO c YES YES WEAK – NO d YES YES WEAK – YES e YES YES STRONG 0.0 NO f YES YES STRONG 0.0 YES FCUP 187 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

g YES YES STRONG 0.5 NO h YES YES STRONG 0.5 YES I YES YES STRONG 1.0 NO j YES YES STRONG 1.0 YES

Fig. 4.1. Frequency of selection of pixels ranked as low (a) and high (b) priorities for CES conservation on the 10 different scenarios, and the absolute number of planning units (PUs(N)) and the planning units weighted by the area of country within the Sahara-Sahel limits (Dinerstein et al., 2017) (PUs (N/country)) selected for CES conservation (c). Pixels (0.5º resolution) are colored according to the number of times they were selected as priorities in the 10 scenarios: less than two times (grey), between two and eight times (yellow), and more than eight times (blue or red). See Fig. D4 and Table 4.2 for scenarios details, and Fig. D1 for country codes.

Further prioritization considerations We performed ranking of CES priorities without generating aggregation across landscape elements (Arponen et al., 2012; Lehtomäki et al., 2009; Moilanen et al., 2005; Moilanen & Wintle, 2007), as it was recently showed that small isolated habitat patches are key for ecological conservation (Wintle et al., 2019) and because some of the features are naturally fragmented

188 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

across the landscape (e.g. rock formations, rock pools and oases; Santarém et al., 2020a). The berm between south-eastern Morocco and north-western Mauritania is heavily mined and militarized (United Nations, 2019). It is impossible to safely visit the region (Santarém et al., 2020b) and most of its biodiversity and ecosystems have been extirpated (Brito et al., 2014). All scenarios were performed excluding it from the analysis, by masking out all the grid cells that intersected with the 2700km-long berm polyline (Kremen et al., 2008; Moilanen, 2013). This procedure ensured that Zonation excluded those pixels since the beginning of the iteration removal process (Moilanen et al., 2014). Our spatial prioritization approach utilized two kinds of data: distribution data of biodiversity features and costs, and structural data elements. The first class of data included high resolution data digitized at the Sahara-Sahel scale, which avoided introducing additional biases in early analysis stages. The second class included mask layers (country borders and the berm boundaries), which are typically known or digitized as polygons with high precision (Pouzols et al., 2014). Additionally, the chosen spatial resolution may had a noteworthy effect on the outcomes of the spatial prioritization (Arponen et al., 2012). However, these potential constraints were likely diluted when using a coarser spatial resolution (pixel size of 0.5º; Brito et al., 2016; Moilanen et al., 2013; Santarém et al., 2020a, 2019; Weyland and Laterra, 2014). Spatial data limitations were, to our best knowledge, halted with all the steps we have taken.

Spatial data analysis: clustering and Maximum Likelihood Classification To account for the spatial variance of the 10 prioritization scenarios for CES availability in Sahara-Sahel, we performed a principal component analysis (PCA) (Brito et al., 2016) (Fig. D5), using the function Principal Components within the Spatial Analyst tool in ArcMap 10.1 (ESRI, 2012). The two PCs were then used to determine the natural grouping (clustering) of the cells in the multidimensional space, using the function Iso Cluster. Following the three classes used in the spatial prioritization outputs (see above) 500 iterations were performed on three classes of the clustering. Given the heteroscedasticity of spatial results, we then estimated the maximum likelihood among the spatial data using the function Maximum Likelihood Classification (MLC). We obtained a raster output that was used to count the number of 0.5° pixels with high availability of CES within each country. We first counted the raw number of 0.5x0.5-degree resolution planning units (PUs(N) hereafter) prioritized for CES and then weighted them by the country-area that falls within the Sahara-Sahel ecoregions (0.5x0.5-degree resolution; PUs(N/country), which will be considered as an index of CES availability hereafter). This procedure ensured that we would account for methodological commission errors related to the FCUP 189 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel spatial prioritization when country borders were considered. For instance, small countries tend to be fully prioritized when compared to large countries because it is relatively easier to protect a large proportion of a small area than of a large area (a full explanation of possible prioritization outcomes from different country sizes can be found in Kukkala et al., 2016; Moilanen et al., 2013; and Pouzols et al., 2014).

Socioeconomic analyses

Indicators of country performance in CES management To be useful for policy decisions, conservation frameworks need to have broad general applicability at global and national levels, and indicators of individualized country-performance in achieving management objectives need to be available, applicable, and representative of the general socioeconomic paradigm of each country (Waldron et al., 2017). We selected a set of the most updated and publicly available variables to represent country performance scores in preserving CES at a national level (Table 4.3): tourism (2 variables: international tourist arrivals - TOU - and ease of movement - VIS); economic (2 variables: gross national income per capita corrected for purchasing-power parity - GNI - and multidimensional poverty - MPI); governance (government effectiveness - GOV - and country’ regulatory environment to conduct business operations - GEF); environment (3 variables: conservation investment levels - CON, biocapacity of each country to regenerate its ecosystems - BIO, and number of threatened species - THR); health (3 variables: tuberculosis - TUB, unsafe water - WAS, human immunodeficiency virus - HIV); social (percentage of population with access to electricity - ELE - and to the internet - INT); and security data (3 variables: peace - GPI, terrorism - GTI, traffic mortalities - ROA). When data were not available for a given year, we took the mean value of that variable from previous available yearly data, which allowed calculating meaningful statistics (Bradshaw & Di Minin, 2019; Daskin & Pringle, 2018). Variables are explained in detail in the Supplementary Material (Text D2).

Statistical analysis The selected indicators are here considered as proxies of the performance of Sahara- Sahel countries in terms of CES management. All indicators were normalized (n-1) before statistical analyses to place effect sizes on a common scale. We performed a PCA (using Pearson’s correlation) of the indicators to identify which variables contribute the most to explain the variability of Sahara-Sahel countries in managing CES. This allowed to empirically quantify

190 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

the influence of socioeconomic variables in the performance of Sahara-Sahel countries in managing CES (Santarém et al., 2018). To avoid interpretation errors due to projection effects common in PCA, and to produce meaningful interpretations for principal components, it is important to identify which variables are associated with the components in question (Peres-Neto et al., 2003). Thus, we used the squared cosines of the socioeconomic normalized variables (Table D1), which reflect the representation quality of a variable on a PC axis and are useful to compare multiple independent variables. Because PC1 explains most of the variance (and the eigenvalue for PC1 was three to four times larger than for PC2 and PC3, respectively), this principal component was used to further evaluate country-performance in managing CES. PC1 was further utilized as an index of country-performance in managing desert CES. Descriptive statistics were summarized prior to performing the PCA (Table D2). All statistical analyses were performed with the statistical software tool XLSTAT (Addinsoft, 2020).

Table 4.3. Categories of the variables used in the statistical analyses, their codes, and data sources. See Text D2 for a description of the variables. The reader is referred to data sources for methodological details behind the data here described.

Category Variables Code Source UNWTO Tourism International tourism, number of arrivals TOU (2019) Passport Passport Index (countries accepting passports VISA-free) VIS Index (2019) World Bank Economic GNI per capita based on purchasing power parity (PPP) GNI (2019) OPHI and Global multidimensional poverty index MPI UNDP (2019) Kaufmann et Governance Government effectiveness GEF al. (2011) World Bank Ease of doing business index BUS (2019) Waldron et Environmental Underfunded countries (conservation spending) CON al. (2013) Global Footprint Biocapacity BIO Network (2019); Lin et al. (2018) Number of threatened species THR IUCN (2019) World Bank Health Prevalence of HIV, total (% of population ages 15-49) HIV (2019) World Bank Incidence of tuberculosis (per 100,000 people) TUB (2019); World Health FCUP 191 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Organization (2019) World Bank (2019); World Mortality rate attributed to unsafe water, unsafe sanitation, and lack of WAS Health hygiene (WASH) (per 100,000 population) Organization (2019) World Bank Social Individuals using the Internet (% of population) INT (2019) World Bank Access to electricity (% of population) ELE (2019) Institute for Economics & Security Global peace index GPI Peace (2019a); Institute for Economics & Global terrorism index GTI Peace (2019b) World Health Mortality caused by road traffic injury (per 100,000 people) ROA Organization (2019)

Measuring countries performance in preserving CES We related the conditions of visiting (i.e., the conditions that each country offers for the usufruct CES) with the availability of CES (i.e., the PUs(N/country) highly ranked for conservation during the prioritization process) in each country. We also compared these conditions to the number of raw planning units with high availability of CES (PUs(N)), the 2017 national visit rates (UNWTO, 2019) (see under “Identification of missing tourism opportunities”), and the 2019 national peacefulness index (Institute for Economics & Peace, 2019). The peace levels of a country are a major concern to tourists visiting regions under conflict (Santarém et al., 2020b). Integration of uncertainty in ecological-economic research has been underlined in recent ecosystem reviews, as it is crucial to inform ecosystem management decisions (e.g. Paul et al., 2020). Thus, we performed a sensitivity test and related the availability of CES with the Human Development Index (HDI), a composite of indicators related to education, life expectancy, wealth and standard of living (UNDP, 2019) that could had been similar to the indicators we collated. We performed additional sensitivity analysis of the potential effect of spatial heterogeneity and prioritization methodological issues, by relating the availability of CES with the PC1 of the PCA (Figs. D8-9) and with the HDI (Figs. D10-11) considering the prioritization ranking results (i.e., with the results obtained before calculating the MLC of the spatial data heteroscedasticity.

192 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Identification of missing tourism opportunities To identify countries that may be missing opportunities to develop tourism activities, we related inbound tourist data (UNWTO, 2019) with the number of PUs (N/Country) with high availability of CES. This allowed to understand which countries are explored to the limit and which ones deserve further exploration, given the number of PUs highly prioritized for CES.

Results

Priorities for cultural services in deserts Our results show that 33.8% of the Sahara-Sahel landscapes display high availability of CES (Fig. 4.2a). Algeria, Chad, Egypt, Libya, Mauritania, Morocco, Niger, and Sudan account for most of the highly ranked areas (i.e., high availability of CES). Yet, when weighting planning units by the area of the country within Sahara-Sahel limits, Burkina Faso, Cameroon, Egypt, Eritrea, Morocco, Senegal, and Tunisia stood out, showing more than half of their country-areas with high availability of CES (Fig. 4.2b). Conducted sensitivity analyses on spatial prioritizations found minor differences according to the methods used (Fig. 4.1) that, nevertheless, do not invalidate the main pattern found. The main mountains and waterbodies of the region were prioritized in both analyses (Figs. 4.1b and 4.2a), although with differences in the number of planning units prioritized within each country (Figs. 4.1c and 4.2b). The Atlantic and Indian coasts also display high availability of CES (Fig. 4.2a). In contrast, the lowest availability of CES is generally located in the largest remote plains of the region (Figs. 4.1a and 4.2a).

FCUP 193 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 4.2. Maximum Likelihood Classification of the two first principal components of the spatial PCA (a), and the absolute number of planning units (PUs(N)) and weighted by the area of country within the Sahara-Sahel limits (PUs (N/country)) highly selected for CES conservation (b). Sahara-Sahel ecoregions (Dinerstein et al., 2017) within Africa are depicted in the small inset. Pixels (0.5º resolution) are colored according to the likelihood of belonging to a given class of CES priority raking: blue – low; yellow – middle; red – high. See Figs. 4.1 and D5 for the frequency of selection of low and high priorities, and for the spatial PCA, respectively, and Fig. D1 for country codes.

Factors conditioning ecosystem services management in Sahara-Sahel The variables with the highest influence in the first axis of the PCA (PC1) were: access to electricity (ELE) and Internet (INT) and tourist arrivals (TOU) with a direct relationship; multidimensional poverty (MPI), governance effectiveness (GEF), ease of doing business (BUS), HIV prevalence (HIV), deaths attributable to unsafe water, sanitation, and hygiene (WAS), and mortality caused by road traffic (ROA) with a inverse relationship. In the second axis of the PCA (PC2) they were: gross national income (GNI), conservation spending (CON), and terrorism (GTI), all with a positive relationship. Variables do not cluster in the PCA, but they are meaningful for this study: nine of them have strong links with the corresponding principal component (see squared cosines in Table D2) and contribute substantially to the first principal component, which account for the highest variance (see eigenvalues and cumulative variance in Table 4.4). Cumulatively, the two first principal components accumulated 61.89% of the variance (PC1: 47.2% and PC2: 14.7%) (Fig. 4.3; Table 4.4) but note that only the PC1 was used as an index to assess country performance in managing CES it explains most of the variance.

194 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 4.3. Results of the Principal Component Analysis to evaluate the potential performance of countries in managing CES. The components that explain most of the variance (PC1 and PC2) are depicted. Countries are represented in blue (see Fig. D1 for country codes) and variables are represented in red (see Table 4.3 for variables codes).

Countries performance in managing cultural ecosystem services Our results show that Algeria, Egypt, Morocco, Senegal, and Tunisia are performing better than all other countries in relation to their capacity to manage CES (Fig. 4.4). In this study, they display better conditions for CES management (PC1 positive values in Fig. 4.4), while also supplying the largest CES in their territories (measured by the widespread availability of CES; Y axis in Fig. 4.4). Among them, Senegal showed higher peacefulness levels than the other countries during the studied period (see symbols in Fig. 4.4) and it stood out as the one with more conditions for visitation and widespread availability of CES. Uncertainty analysis to development conditions revealed that Senegal displays distinct conditions according to the indexes used: whereas the index based in our multivariate analysis of socioeconomic indicators revealed that the country provides more conditions for the usufruct of CES (Fig. 4.4), the Human Development Index (HDI) positioned the country to the group providing worse conditions (Figs. 4.5 and D7).

FCUP 195 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table 4.4. Loadings of the principal components (PCs), the eigenvalues (EIG), the variance (VAR), and the cumulative variance (CUM) of the principal components to evaluate the potential performance of countries in managing CES supply. Predictors highly correlated with the two first principal components (that cumulative explain most of the variance) are marked in bold. For simplicity, only the first six PCs are shown. See Table 4.3 for socio-economic variable codes and Table D2 for the representation quality (squared cosines) of the variables each PC.

PC1 PC2 PC3 PC4 PC5 PC6 TOU 0.779 -0.006 0.369 0.101 -0.223 -0.252 VIS 0.625 -0.417 0.349 0.248 -0.215 0.270 GNI 0.630 0.667 -0.098 -0.130 -0.181 0.246 MPI -0.781 -0.386 -0.142 0.410 0.075 -0.063 GEF 0.796 -0.361 0.229 0.208 0.152 0.185 BUS -0.796 0.294 -0.339 -0.339 0.078 0.103 CON -0.386 0.600 0.258 0.439 0.010 0.339 BIO -0.653 -0.419 0.362 -0.295 -0.336 -0.042 THR 0.065 0.208 0.625 -0.267 0.674 -0.103 HIV -0.711 0.070 0.620 -0.120 0.109 -0.012 TUB -0.607 -0.106 0.625 -0.204 -0.327 0.107 WAS -0.778 -0.260 0.222 0.223 0.039 -0.225 INT 0.901 0.063 0.293 -0.042 -0.043 0.022 ELE 0.873 0.280 0.238 -0.226 -0.057 -0.040 GPI -0.647 0.581 0.042 -0.059 -0.348 -0.189 GTI -0.379 0.642 0.295 0.500 0.008 -0.132 ROA -0.744 -0.150 0.070 -0.106 0.083 0.547

EIG 8.020 2.501 2.102 1.188 0.984 0.791 VAR (%) 47.18 14.71 12.36 6.99 5.79 4.65 CUM (%) 47.18 61.89 74.25 81.24 87.03 91.68

196 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 4.4. Relationship between the conditions for visitation and the availability of CES in each country. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socioeconomic indicators are constraining national conditions for CES (Table 4.4 and Fig. 4.3). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country)). Country symbols are colored according to the country-area highly ranked for CES conservation and the 2017 tourism inbounds (UNWTO, 2019): yellow (tourism over-explored); red (tourism under-explored); and black (tourism regularly explored). Symbol shapes represent countries-rankings according to the 2019 index of state of peace (Institute for Economics & Peace, 2019): square – high; circles – medium; diamonds – low; and triangles – very low. Symbol sizes are proportional to the raw number of planning units (PUs(N)) highly ranked for CES (see Fig. D6). See Fig. D1 for country codes.

FCUP 197 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 4.5. Sensitivity analysis of the relationship between the conditions for visitation and the availability of CES in each country. Conditions for visitation were retrieved from the Human Development Index (HDI) (UNDP, 2019). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country). Country symbols are coded according to the country-area (YY axis) highly ranked for CES conservation and the 2017 tourism inbounds (UNWTO, 2019): yellow (tourism over-explored); red (tourism under-explored); and black (tourism regularly explored). Symbol shapes represent countries-rankings according to the 2019 index of state of peace (Institute for Economics & Peace, 2019a): square – high; circles – medium; diamonds – low; and triangles – very low. Symbol sizes represent the raw number of planning units – PUs(N) – highly ranked for CES conservation (see Fig. D7). Thresholds of medium developed countries (0.55) and high developed countries (0.70) are marked with a dashed line. See Fig. D1 for country codes.

Countries missing opportunities for local development We showed that Cameroon, Eritrea, and Senegal are missing opportunities to develop their social and economic conditions based on sustainable ecotourism (red symbols in Figs. 4.4 and 4.6), whereas Egypt, Morocco, and Tunisia have already explored their CES resources to the limit or even beyond (yellow symbols in Figs. 4.4 and 4.6). Most of the other analyzed countries display low availability of CES and less favorable conditions for ecotourism visitation (black dots in Figs. 4.4 and 4.6).

198 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 4.6. Missing opportunities for local development among Sahara-Sahel countries. Missing opportunities are defined according to the number of planning units weighted by the country-area within the Sahara-Sahel ecoregions (Dinerstein et al., 2017) (PUs (N/country)) prioritized for CES conservation and the inbound tourists in each country. Country symbols are colored as: red (tourism under-explored), yellow (tourism explored to the limit), and black (tourism normally explored when compared to the number of PUs identified with high availability of CES). Tourism data (UNWTO, 2019) are in millions. See Fig. D1 for country codes.

Discussion

We have made use of several sources of spatial and socioeconomic data, including high- resolution indicators of constraints to CES management. This is the most complete existing spatial and analytical study ever conducted in desert ecosystems. To meet our study objectives, it was crucial to account for several spatial-explicit species, landscape, and cultural data, as well as constraints to CES (Table 4.1), and three simulations with 10 different scenarios (Table 4.2).

Prioritization of CES in Sahara-Sahel landscapes The main mountains and wetlands of the Sahara-Sahel, and the Atlantic and Indian coasts display high availability of CES (Figs. 4.1b and 4.2a). These regions have been previously identified as hotspots of CES (Santarém et al., 2020b) and of flagship-species for conservation and ecotourism promotion (Santarém et al., 2019), highlighting their importance for ecosystem- FCUP 199 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel based conservation (Mengist et al., 2020). Yet, these regions have attracted very little scientific attention (Durant et al., 2014). In contrast, the lowest availability of CES were generally located in the largest remote plains of the region (Figs. 4.1a and 4.2a), where armed conflicts and overexploitation of ecosystems are widespread, with substantial impacts on people and on the services they could had benefited from if geopolitical tensions were ameliorated (Brito et al., 2014, 2018; Santarém et al., 2020b).

Priorities for CES management were found to be high in Burkina Faso, Cameroon, Egypt, Eritrea, Morocco, Senegal, and Tunisia. Yet, Algeria, Chad, Egypt, Mauritania, Mali, Niger, and Morocco showed a higher number of raw planning units (PUs(N)) highly ranked for CES management. These differences are related to countries’ surface areas: while the latter group of countries are the largest in Sahara-Sahel, the former group of countries have limited areas within the Sahara-Sahel ecoregions. Thus, for example, Algeria has larger numbers of areas providing CES, but its substantial size diminishes its overall CES availability at the regional level, which may condition results interpretation (Moilanen et al., 2013, 2014). These size effects and corresponding prioritization performance are discussed in the literature (see, for instance, Kukkala et al., 2016; Moilanen et al., 2013; Moilanen & Arponen, 2011). Yet, by accounting for both the number of raw planning units (PUs(N)) and the planning units weighted by the country-area that falls within the Sahara-Sahel ecoregions (PUs(N/country) we have balanced priorities with regional to global preferences in the spatial prioritization process (Moilanen & Arponen, 2011).

Countries performance in managing Cultural Ecosystem Services

Algeria, Egypt, Morocco, Senegal, and Tunisia are performing better than all other countries in relation to their capacity to manage CES. They not only showed higher availability of CES (except Algeria; PUs (N/country) = 0.3), but also more conditions for visitation. Among these countries, Senegal showed higher peacefulness levels than the other countries during the studied period. Sensitivity analysis to development conditions revealed that Senegal displays distinct conditions according to the indexes used. Note that HDI includes indicators such as “Life expectancy at birth”, “Expected years of schooling”, “Mean years of schooling”, and “Gross national income per capita” whereas our index goes further and includes many more components of different areas (see Table 4.3). Rather than Senegal, all other Sahara-Sahel countries need to improve safety conditions to attract international tourism, or else the consequences of long-term conflict will be nefarious for local communities seeking sustainability (Ospina, 2006). Cameroon, Chad, Mali, Niger and Nigeria are particularly vulnerable to terrorism and conflicts, and security

200 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

in Burkina Faso, Egypt, Libya, Sudan and South Sudan needs to be reinforced (OECD/SWAC, 2014; Walther, 2017). Ameliorating conflicts and improving hospitality conditions in these countries will enable the attraction of ecotourists, which potentially can contribute to the sustainable development of local societies (Santarém et al., 2020a).

Biodiversity losses related to weak governance and socioeconomic factors may accelerate losses in ecosystem functioning (Cardinale et al., 2012), threatening nature services to people (Blicharska et al., 2019; Santarém et al., 2020b). Looking forward, the picture for low-income countries is much more uncertain than for high-income countries, in general, as bad governance and poor socioeconomic conditions may limit plans for ecosystem services conservation. Globally, countries with weak governance and high political corruption are associated with poor conservation performance (Bradshaw & Di Minin, 2019) and lower visitation levels (Hausmann, Toivonen, Heikinheimo, et al., 2017), and the same pattern is found along Sahara-Sahel countries (Brito et al., 2018). For instance, a large proportion of Niger’s and Chad’s natural ecosystems remain intact (Di Marco et al., 2019) and offer prospects for CES, yet they need to improve levels of social organization, stability and good governance to achieve the goals of conserving CES (Maron et al., 2019). Niger has committed to conservation programs targeting key biodiversity (Durant et al., 2014), but recently the Government decided to declassify part of the Termit and Tin-Toumma National Nature Reserve (RNNTT) in favor of oil exploitation activities, threatening whole ecosystems that could supply CES to people if remained protected (IUCN SSC Antelope Specialist Group, 2020). Obviously, also health and safety issues need to be improved if countries want to increase ecotourism based on their CES. Severe health issues keep untreated along Sahelian countries (Beasley et al., 2002; Yakum et al., 2017), furthering a human crisis that fail to cease. New opportunities for tourism can be developed if basic needs are met in the short/medium term.

Our analyses showed that access to electricity and Internet are the most important factors in conditioning the performance of Sahara-Sahel countries in managing CES, as it is the case across all Africa (Hausmann, Toivonen, Heikinheimo, et al., 2017). Combined with the other strongest indicators of countries performance in managing CES, our analyses revealed that CES in Sahara-Sahel are conditioned by many contextual human and political factors that deserve thoroughly consideration when developing national and especially supra-national conservation plans that are much needed in the Sahara-Sahel region (Brito et al., 2016). We contend the fact that we used indicators from many different fields: tourism, socioeconomic, governance, environmental, health, and security (Table 4.3). Individual countries, however, can be analyzed FCUP 201 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel based on a smaller group of indicators when performing national level identification of optimum or near-optimum sites for CES provision and demand.

Missing opportunities for sustainable development

Countries such as Cameroon, Eritrea, and Senegal are missing opportunities for sustainable development. They need to improve their socioeconomic conditions to attract more international ecotourists to benefit from desert CES and economically contribute to the sustainable development of the region. Particularly, Eritrea is among the poorest and least visited countries in the world (UNWTO, 2019), but showed a considerable availability of CES in this study. In contrast, Egypt, Morocco, and Tunisia are receiving far more tourists to benefit from their CES than what their ecosystems possess (CES availability between 0.52 and 0.63). Until 2017, these countries received 235 to 378 times more tourists than the poorest and least visited countries in Sahara-Sahel (e.g., Chad; Fig. D2). This disparity between country’ tourism arrivals calls for urgent balancing at the regional level, as otherwise there is a serious risk of overexploiting some types of ecosystems while others remain underexplored (OECD/SWAC, 2014). Although most of the other countries are not missing nor overexploiting opportunities, they could develop strategies that will enable them to attract more international ecotourists and improve their socioeconomic conditions. For instance, some of the rarest charismatic desert species (e.g. Nanger dama and Addax nasomaculatus) still persist within their territories and could be used for synergic ecotourism and conservation promotion (Durant et al., 2014; Santarém et al., 2019).

Potential data limitations and methodological constraints

Our work is based on the availability of big datasets containing georeferenced data on features and costs for the prioritization of natural resources in the Sahara-Sahelian region integrated with country-level socio-economic data to understand countries-performance in managing their CES. While using big datasets presents advantages related with data availability, it also presents limitations related with data quality and local-scale processes, which may have conditioned the analyses performed (Cerretelli et al., 2018). The lack of high-resolution data for some regions in the Sahara-Sahel has been identified in previous research (Brito et al., 2014, 2016), and the use of a coarse spatial resolution may have compromise, for instance, the efficiency of prioritization rankings. Still, a large spatial resolution is needed given the large extent of Sahara-Sahel and the computational power to perform the spatial prioritization demands these compromises (Brito et al., 2016). Even if this approach may limit local-scale interpretations in

202 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

some parts of the region, the research here developed offers hints to develop CES planning at the regional level that, nevertheless, can be adapted to finer-scale levels once future detailed data is available.

Social media data can potentially be used for collecting meaningful ecological and ecotourism data in the future (Hausmann et al., 2019; Hausmann, Toivonen, Heikinheimo, et al., 2017; Hausmann, Toivonen, Slotow, et al., 2017; Toivonen et al., 2019). But such meaningful data will only be possible to collect once localized conflict events are ameliorated (Brito et al., 2018), which can then lead to increased ecotourist visitation levels in Sahara-Sahel that allow assembling these data (Santarém et al., 2020b). Other constraints, such as natural resources extraction facilities or smuggling/trafficking routes that threaten ecosystems and unable people to benefit from natural resources (Brito et al., 2018; Levin et al., 2019; Taylor et al., 2017) can be considered in future studies to highlight the costs related to CES losses. But these elements are highly dynamic in space and time, making them hard to include in CES studies (Santarém et al., 2020a).

Our quantitative approach for evaluating CES is based in a specific index developed for such purpose that uses socio-economic variables as proxies of the performance of Sahara-Sahel countries in terms of CES management. There are other alternative approaches, such as the HDI composite index, which comprises indicators related to education, life expectancy, wealth and standard of living (UNDP, 2019). However, our index goes further in considering additionally tourism, environmental, governance and security indicators, which further details the regional development situation of each Sahara-Sahel country, and thus provides a clearer picture of CES assessment and management.

Our spatial approach is based in the most recent publicly available decision-support tool to produce rankings for the conservation and management of biodiversity (Lehtomäki & Moilanen, 2013; Moilanen et al., 2014), which enables the production of prioritized maps reflecting the rank order of importance of planning units (Moilanen, 2007; Moilanen et al., 2005, 2014). There are numerous alternative methods to evaluate ecosystem services (see Brown & Fagerholm, 2015; Crossman et al., 2013; Kosanic & Petzold, 2020; Martnez-Harms & Balvanera, 2012; Milcu et al., 2013; Wolff et al., 2015 for global reviews on CES methods), including biophysical modelling, simple GIS mapping, or methods focused on participatory methods to evaluate people preferences (Harrison et al., 2018). However, in our case, we opted to use decision-support tools has they are useful for conservation and management research, despite their application to ecosystem research is just start growing (see for example Di Minin et al., 2017). Furthermore, by FCUP 203 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel applying a novel mixed-method approach, we were able to assess the performance of different states in the conservation of and benefits from CES, something never done before globally.

The PCA developed in this study is unbalanced between the number of countries (18) and the number of socio-economic variables (17) under study. However, the PCA is a statistical approach that assumes that the number of variables should be many times larger than the number of countries to find meaningful patterns (Björklund, 2019). We addressed this issue by testing the distinctness of the different PCs and used only the most distinct one which displayed the eigenvalue three to four times larger (8.02) than the next two PCs (PC2: 2.5 and PC3 2.1; Table 4.4). We also presented the squared cosines (Table D2), which reflect the representation quality of the variables on the different PCs and avoided interpretation errors due to projections (Peres- Neto et al., 2003). These steps were taken to minimize possible biases in PCA results.

Future research also needs to refine methods to account for complex and intertwining temporal dynamics of conflicts (Hanaček & Rodríguez-Labajos, 2018), which can rank countries’ performance differently along time. For instance, although the peace situation of Burkina Faso was ranked as medium during the studied period, the very recent escalating conflict in the country (in early 2020) most probably influenced the safety perceptions of potential visitors. As such, the changing dynamics of conflicts need to be considered carefully in regional planning, as conditions for visitation can quickly deteriorate or ameliorate.

Recommendations for desert CES management

Globally, ecosystem services derive values to people in the order of USD145trillion/year, a figure that is 4.5 times the value of 2014 Gross World Product (Costanza et al., 2014). Although CES remain poorly understood and neglected (Cheng et al., 2019), their estimated economic contribution to people’s mental health alone yields around 2019 USD6 trillion (Buckley et al., 2019), which represents about 8% of the 2017 Global National Product. These are very high figures concerning the services nature provides to people. Deserts’ contributions to these figures may be unspotted (Taylor et al., 2017), especially the economic value of desert CES. Studies estimating the value of desert CES are desperately needed to understand their economic and social contribution (Durant et al., 2012; Santarém et al., 2020b), particularly to countries such as Chad, Central African Republic, Eritrea, Ethiopia, Nigeria, Mali, and South Sudan, which display the lowest economic conditions among the studied countries and where the recognized value of CES could significantly improve their economies. For example, the annual value of CES in the

204 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Big Bend region of the Chihuahuan Desert was estimated to be around 2015 USD37.82 per hectare, a very high number given the threats that the region is under (Taylor et al., 2017). Nevertheless, we call countries and regional developers to opt for desert ecotourism activities that suit their particular ecological and sociocultural contexts (Santarém et al., 2020a; Santarém & Paiva, 2015; UNESCO, 2003b), in contrast to the forms of mainstream tourism that focus on growth instead of development and that leads to negative impacts on local ecosystems and communities (Buckley, 2011).

We call the regional and international community to jointly consider setting the following initiatives, which, although being quite ambitious, will be essential to deal with nature losses while developing local economies: 1) perform valuations of the natural capital to highlight the economic importance of desert ecosystems to people (Buckley et al., 2019; Costanza et al., 2014; de Groot et al., 2012; Strand et al., 2018; Teoh et al., 2019); 2) split conservation budgets to expand the land under protection (Pouzols et al., 2014), particularly by developing transboundary conservation areas (Brito et al., 2016) across hotspots of biodiversity (Brito et al., 2016; Santarém et al., 2019) and of ecosystem services (Santarém et al., 2020b), and to properly manage existing protected areas (Adams et al., 2019) to ensure the multifunctioning of ecosystems at multiple places and times (Cardinale et al., 2012); 3) strengthen environmental laws (Willemen et al., 2020) through international accepted Arms Trade Treaties to control arms trafficking (Brito et al., 2018) and to diminish armed conflicts in endangered World Heritage Sites of substantial global value (Levin et al., 2019); 4) improve governance, transparency and accountability (Díaz et al., 2019; Maron et al., 2019) to enable local societies to be more open to businesses (Brito et al., 2018); 5) promote community-based management of natural resources (Brito et al., 2018; Garnett et al., 2018) to ensure that their expertise is considered in conservation plans (Tilman et al., 2017; Willemen et al., 2020); and 6) develop community-based and flagship-based ecotourism to improve local social and economic conditions and to preserve threatened biodiversity in the long run (Brito et al., 2018; Santarém et al., 2019; UNEP, 2006b).

At a time when the outlook for nature conservation seems bleak, this study showed that promoting sustainable development practices to conserve CES in deserts will allow the proper functioning of ecosystems and their capacity to provide society with the essential services needed to prosper (Blicharska et al., 2019; Cardinale et al., 2012). The framework developed here presents a constructive opportunity for shaping global policies towards ecosystem management and for highlighting the contributions of cultural services to the SDGs (Wood et al., 2018).

FCUP 205 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Conclusion

Here, an interdisciplinary approach was employed to understand the performance of individual countries in managing their desert CES. The work showed that not only desert ecosystems provide many benefits to people but that there are opportunities for sustainable local development that are not fully explored. The results have implications for regional planners and policymakers that need to manage local ecosystem services to attract more ecotourists able to contribute to the sustainable development of the region. Still, national differences need to be accounted when developing regional plans for ecosystem management. Any approach to manage ecosystems in the Sahara-Sahel needs to be further contextualized and underline country’ specificities, as emphasized in this paper. Combining decision support tools with socioeconomic data provided a first step to optimally allocate investments for conservation management. This is especially relevant in a region with big development contrasts, but where most of the countries are still underdeveloped despite displaying large cultural services that can benefit the global society. Policy makers can use the framework provided here to achieve regional targets of the SDGs. The framework is replicable to other areas where regional planning is needed for CES management and where country differences need to be taken in consideration.

206 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Acknowledgements Financial support was given by AGRIGEN-NORTE-01-0145-FEDER- 000007, supported by Norte Portugal Regional Operational Programme (NORTE2020), under the PORTUGAL 2020 Partnership Agreement, through the European Regional Development Fund (ERDF). F.S. and J.C.B. are supported by Fundação para a Ciência e Tecnologia through grant PD/BD/132407 and contract CEECINST/00014/2018; J.S. is supported by Academy of Finland through grant 272168. We thank BirdLife International. IUCN, the World Bank, and the United Nations for giving free access to public data used in this work, and Sílvia Carvalho for early discussions about Zonation software. We thank the comments made by the editor Dr. Paulo Pereira and the two anonymous reviewers, which improved the manuscript. FCUP 207 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

References Adams, V. M., Iacona, G. D., & Possingham, H. P. (2019). Weighing the benefits of expanding protected areas versus managing existing ones. Nature Sustainability, 2(5), 404–411. http://dx.doi.org/10.1038/s41893-019-0275-5

Addinsoft. (2020). XLSTAT 2020.1.1: Data Analysis and Statistics Software for Microsoft Excel (2020.1.1). Addinsoft.

Arponen, A., Cabeza, M., Eklund, J., Kujala, H., & Lehtomäki, J. (2010). Costs of integrating economics and conservation planning. Conservation Biology, 24(5), 1198–1204. https://doi.org/10.1111/j.1523-1739.2010.01539.x

Arponen, A., Heikkinen, R. K., Thomas, C. D., & Moilanen, A. (2005). The value of biodiversity in reserve selection: Representation, species weighting, and benefit functions. Conservation Biology, 19(6), 2009–2014. https://doi.org/10.1111/j.1523-1739.2005.00218.x

Arponen, A., Lehtomäki, J., Leppänen, J., Tomppo, E., & Moilanen, A. (2012). Effects of connectivity and spatial resolution of analyses on conservation prioritization across large extents. Conservation Biology, 26(2), 294–304. https://doi.org/10.1111/j.1523-1739.2011.01814.x

Bachi, L., Ribeiro, S. C., Hermes, J., & Saadi, A. (2020). Cultural Ecosystem Services (CES) in landscapes with a tourist vocation: Mapping and modeling the physical landscape components that bring benefits to people in a mountain tourist destination in southeastern Brazil. Tourism Management, 77(May 2019). https://doi.org/10.1016/j.tourman.2019.104017

Beasley, M., Brooker, S., Ndinaromtan, M., Madjiouroum, E. M., Baboguel, M., Djenguinabe, E., & Bundy, D. A. P. (2002). First nationwide survey of the health of schoolchildren in Chad. Tropical Medicine and International Health, 7(7), 625–630. https://doi.org/10.1046/j.1365- 3156.2002.00900.x

Björklund, M. (2019). Be careful with your principal components. Evolution, 73(10), 2151–2158. https://doi.org/10.1111/evo.13835

Blicharska, M., Smithers, R. J., Mikusiński, G., Rönnbäck, P., Harrison, P. A., Nilsson, M., & Sutherland, W. J. (2019). Biodiversity’s contributions to sustainable development. Nature Sustainability, 2(12), 1083–1093. https://doi.org/10.1038/s41893-019-0417-9

208 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Bradshaw, C. J. A., & Di Minin, E. (2019). Socio-economic predictors of environmental performance among African nations. Scientific Reports, 9(9306), 1–13. https://doi.org/10.1038/s41598-019- 45762-3

Brito, J. C., Durant, S. M., Pettorelli, N., Newby, J., Canney, S., Algadafi, W., Rabeil, T., Crochet, P.- A., Pleguezuelos, J. M., Wacher, T., Smet, K. De, Gonçalves, D. V., da Silva, M. J. F., Martínez- Freiría, F., Abáigar, T., Campos, J. C., Comizzoli, P., Fahd, S., Fellous, A., … Carvalho, S. B. (2018). Armed conflicts and wildlife decline : Challenges and recommendations for effective conservation policy in the Sahara-Sahel. Conservation Letters, 11, 1–13. https://doi.org/10.1111/conl.12446

Brito, J. C., Godinho, R., Martínez-Freiría, F., Pleguezuelos, J. M., Rebelo, H., Santos, X., Vale, C. G., Velo-Antón, G., Boratyński, Z., Carvalho, S. B., Ferreira, S., Gonçalves, D. V., Silva, T. L., Tarroso, P., Campos, J. C., Leite, J. V., Nogueira, J., Alvares, F., Sillero, N., … Carranza, S. (2014). Unravelling biodiversity, evolution and threats to conservation in the sahara-sahel. Biological Reviews, 89(1), 215–231. https://doi.org/10.1111/brv.12049

Brito, J. C., Tarroso, P., Vale, C. G., Martínez-Freiría, F., Boratyński, Z., Campos, J. C., Ferreira, S., Godinho, R., Gonçalves, D. V., Leite, J. V., Lima, V. O., Pereira, P., Santos, X., da Silva, M. J. F. . F., Silva, T. L., Velo-Antón, G., Veríssimo, J., Crochet, P. A., Pleguezuelos, J. M., & Carvalho, S. B. (2016). Conservation biogeography of the Sahara-Sahel: Additional protected areas are needed to secure unique biodiversity. Diversity and Distributions, 22(4), 371–384. https://doi.org/10.1111/ddi.12416

Brown, G., & Fagerholm, N. (2015). Empirical PPGIS/PGIS mapping of ecosystem services: A review and evaluation. Ecosystem Services, 13, 119–133. https://doi.org/10.1016/j.ecoser.2014.10.007

Buckley, R. (2011). Tourism and Environment. Annu. Rev. Environ. Resour, 36, 397–416. https://doi.org/10.1146/annurev-environ-041210-132637

Buckley, R., Brough, P., Hague, L., Chauvenet, A., Fleming, C., Roche, E., Sofija, E., & Harris, N. (2019). Economic value of protected areas via visitor mental health. Nature Communications, 10(1), 1–10. https://doi.org/10.1038/s41467-019-12631-6

Burkhard, B., Kroll, F., Nedkov, S., & Müller, F. (2012). Mapping ecosystem service supply, demand and budgets. Ecological Indicators, 21, 17–29. https://doi.org/10.1016/j.ecolind.2011.06.019 FCUP 209 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Cardinale, B. J., Duffy, J. E., Gonzalez, A., Hooper, D. U., Perrings, C., Venail, P., Narwani, A., MacE, G. M., Tilman, D., Wardle, D. A., Kinzig, A. P., Daily, G. C., Loreau, M., Grace, J. B., Larigauderie, A., Srivastava, D. S., & Naeem, S. (2012). Biodiversity loss and its impact on humanity. Nature, 486(7401), 59–67. https://doi.org/10.1038/nature11148

Casado-Arzuaga, I., Onaindia, M., Madariaga, I., & Verburg, P. H. (2014). Mapping recreation and aesthetic value of ecosystems in the Bilbao Metropolitan Greenbelt (northern Spain) to support landscape planning. Landscape Ecology, 29(8), 1393–1405. https://doi.org/10.1007/s10980-013- 9945-2

Cerretelli, S., Poggio, L., Gimona, A., Yakob, G., Boke, S., Habte, M., Coull, M., Peressotti, A., & Black, H. (2018). Spatial assessment of land degradation through key ecosystem services: The role of globally available data. Science of the Total Environment, 628–629, 539–555. https://doi.org/10.1016/j.scitotenv.2018.02.085

Cheng, X., Van Damme, S., Li, L., & Uyttenhove, P. (2019). Evaluation of cultural ecosystem services: A review of methods. Ecosystem Services, 37, 100925. https://doi.org/10.1016/j.ecoser.2019.100925

Costanza, R., de Groot, R., Sutton, P., van der Ploeg, S., Anderson, S. J., Kubiszewski, I., Farber, S., & Turner, R. K. (2014). Changes in the global value of ecosystem services. Global Environmental Change, 26(1), 152–158. https://doi.org/10.1016/j.gloenvcha.2014.04.002

Crossman, N. D., Burkhard, B., Nedkov, S., Willemen, L., Petz, K., Palomo, I., Drakou, E. G., Martín- Lopez, B., McPhearson, T., Boyanova, K., Alkemade, R., Egoh, B., Dunbar, M. B., & Maes, J. (2013). A blueprint for mapping and modelling ecosystem services. Ecosystem Services, 4, 4– 14. https://doi.org/10.1016/j.ecoser.2013.02.001

Daskin, J. H., & Pringle, R. M. (2018). Warfare and wildlife declines in Africa’s protected areas. Nature, 553(7688), 328–332. https://doi.org/10.1038/nature25194 de Groot, R., Brander, L., van der Ploeg, S., Costanza, R., Bernard, F., Braat, L., Christie, M., Crossman, N., Ghermandi, A., Hein, L., Hussain, S., Kumar, P., McVittie, A., Portela, R., Rodriguez, L. C., ten Brink, P., & van Beukering, P. (2012). Global estimates of the value of ecosystems and their services in monetary units. Ecosystem Services, 1(1), 50–61. https://doi.org/10.1016/j.ecoser.2012.07.005

210 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Di Marco, M., Ferrier, S., Harwood, T. D., Hoskins, A. J., & Watson, J. E. M. (2019). Wilderness areas halve the extinction risk of terrestrial biodiversity. Nature, 573, 582–585. https://doi.org/https://doi.org/10.1038/s41586-019-1567-7

Di Minin, E., & Moilanen, A. (2012). Empirical evidence for reduced protection levels across biodiversity features from target-based conservation planning. Biological Conservation, 153, 187–191. https://doi.org/10.1016/j.biocon.2012.04.015

Di Minin, E., Soutullo, A., Bartesaghi, L., Rios, M., Szephegyi, M. N., & Moilanen, A. (2017). Integrating biodiversity, ecosystem services and socio-economic data to identify priority areas and landowners for conservation actions at the national scale. Biological Conservation, 206, 56–64. https://doi.org/10.1016/j.biocon.2016.11.037

Díaz, S., Pascual, U., Stenseke, M., Martín-López, B., Watson, R. T., Molnár, Z., Hill, R., Chan, K. M. A., Baste, I. A., Brauman, K. A., Polasky, S., Church, A., Lonsdale, M., Larigauderie, A., Leadley, P. W., van Oudenhoven, A. P. E., van der Plaat, F., Schröter, M., Lavorel, S., … Shirayama, Y. (2018). Assessing nature’s contributions to people. Science, 359(6373), 270–272. https://doi.org/10.1126/science.aap8826

Díaz, S., Settele, J., Brondízio, E. S., Ngo, H. T., Agard, J., Arneth, A., Balvanera, P., Brauman, K. A., Butchart, S. H. M., Chan, K. M. A., Lucas, A. G., Ichii, K., Liu, J., Subramanian, S. M., Midgley, G. F., Miloslavich, P., Molnár, Z., Obura, D., Pfaff, A., … Zayas, C. N. (2019). Pervasive human- driven decline of life on Earth points to the need for transformative change. Science, 366(6471), eaax3100. https://doi.org/10.1126/science.aax3100

Dinerstein, E., Olson, D., Joshi, A., Vynne, C., Burgess, N. D., Wikramanayake, E., Hahn, N., Palminteri, S., Hedao, P., Noss, R., Hansen, M., Locke, H., Ellis, E. C., Jones, B., Barber, C. V., Hayes, R., Kormos, C., Martin, V., Crist, E., … Kindt, R. (2017). An ecoregion-based approach to protecting half the terrestrial realm. BioScience, 67(6), 534–545. https://doi.org/10.1093/biosci/bix014

Dioko, L. (Don) A. N., & Harrill, R. (2019). Killed while traveling – Trends in tourism-related mortality, injuries, and leading causes of tourist deaths from published English news reports, 2000–2017 (1H). Tourism Management, 70(August 2018), 103–123. https://doi.org/10.1016/j.tourman.2018.08.002 FCUP 211 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Durant, S. M., Pettorelli, N., Bashir, S., Woodroffe, R., Wacher, T., Ornellas, P. de, Ransom, C., Abdelgadir, M., Alqamy, H. E. L., Beddiaf, M., Belbachir, F., Berbash, A., Boitani, L., Breitenmoser, C., Cano, M., Chardonnet, P., Collen, B., Cornforth, W., Cuzin, F., … Baillie, J. (2012). Forgotten biodiversity in desert ecosystems. Science, 336(6087), 1379–1380. https://doi.org/10.1126/science.336.6087.1379

Durant, S. M., Wacher, T., Bashir, S., Woodroffe, R., De Ornellas, P., Ransom, C., Newby, J., Abáigar, T., Abdelgadir, M., El Alqamy, H., Baillie, J., Beddiaf, M., Belbachir, F., Belbachir-Bazi, A., Berbash, A. A., Bemadjim, N. E., Beudels-Jamar, R., Boitani, L., Breitenmoser, C., … Pettorelli, N. (2014). Fiddling in biodiversity hotspots while deserts burn? Collapse of the Sahara’s megafauna. Diversity and Distributions, 20(1), 114–122. https://doi.org/10.1111/ddi.12157

ESRI. (2012). ArcGIS Release 10.1. Redlands.

GADM. (2018). Global Administrative Areas v3.6. https://gadm.org/data.html

Garnett, S. T., Burgess, N. D., Fa, J. E., Fernández-Llamazares, Á., Molnár, Z., Robinson, C. J., Watson, J. E. M., Zander, K. K., Austin, B., Brondizio, E. S., Collier, N. F., Duncan, T., Ellis, E., Geyle, H., Jackson, M. V., Jonas, H., Malmer, P., McGowan, B., Sivongxay, A., & Leiper, I. (2018). A spatial overview of the global importance of Indigenous lands for conservation. Nature Sustainability, 1(7), 369–374. https://doi.org/10.1038/s41893-018-0100-6

Hanaček, K., & Rodríguez-Labajos, B. (2018). Impacts of land-use and management changes on cultural agroecosystem services and environmental conflicts—A global review. Global Environmental Change, 50(February), 41–59. https://doi.org/10.1016/j.gloenvcha.2018.02.016

Harrison, P. A., Dunford, R., Barton, D. N., Kelemen, E., Martín-López, B., Norton, L., Termansen, M., Saarikoski, H., Hendriks, K., Gómez-Baggethun, E., Czúcz, B., García-Llorente, M., Howard, D., Jacobs, S., Karlsen, M., Kopperoinen, L., Madsen, A., Rusch, G., van Eupen, M., … Zulian, G. (2018). Selecting methods for ecosystem service assessment: A decision tree approach. Ecosystem Services, 29, 481–498. https://doi.org/10.1016/j.ecoser.2017.09.016

Hausmann, A., Toivonen, T., Fink, C., Heikinheimo, V., Tenkanen, H., Butchart, S. H. M., Brooks, T. M., & Di Minin, E. (2019). Assessing global popularity and threats to Important Bird and Biodiversity Areas using social media data. Science of the Total Environment, 683, 617–623. https://doi.org/10.1016/j.scitotenv.2019.05.268

212 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Hausmann, A., Toivonen, T., Heikinheimo, V., & Tenkanen, H. (2017). Social media reveal that charismatic species are not the main attractor of ecotourists to sub-Saharan protected areas. Scientific Reports, March, 1–9. http://dx.doi.org/10.1038/s41598-017-00858-6

Hausmann, A., Toivonen, T., Slotow, R., Tenkanen, H., Moilanen, A., Heikinheimo, V., & Minin, E. Di. (2017). Social media data can be used to understand tourists’ preferences for nature-based experiences in protected areas. Conservation Letters, 1–10.

He, S., Su, Y., Shahtahmassebi, A. R., Huang, L., Zhou, M., Gan, M., Deng, J., Zhao, G., & Wang, K. (2019). Assessing and mapping cultural ecosystem services supply, demand and flow of farmlands in the Hangzhou metropolitan area, China. Science of the Total Environment, 692, 756–768. https://doi.org/10.1016/j.scitotenv.2019.07.160

Hosni, E. (2000). Strategy for Sustainable Tourism Development in the Sahara. In UNESCO (Issue 32). http://www.unesco.org/new/en/culture/themes/dynamic-content-single- view/news/strategy_for_sustainable_tourism_development_in_the_sahara/

Institute for Economics & Peace. (2019). Global Peace Index 2019: Measuring Peace in a Complex World. http://visionofhumanity.org/reports

IUCN SSC Antelope Specialist Group. (2020). IUCN mission to Niger for the conservation of the last wild addax and dama gazelles and the Termit and Tin Toumma National Nature Reserve : Report and Recommendations. https://www.iucn.org/ssc-groups/mammals/mammals-a- e/antelope/resources

Kosanic, A., & Petzold, J. (2020). A systematic review of cultural ecosystem services and human wellbeing. Ecosystem Services, 45(October 2019), 101168. https://doi.org/10.1016/j.ecoser.2020.101168

Kremen, C., Cameron, A., Moilanen, A., Phillips, S. J., Thomas, C. D., Beentje, H., Dransfield, J., Fisher, B. L., Glaw, F., Good, T. C., Harper, G. J., Hijmans, R. J., Lees, D. C., Louis, E., Nussbaum, R. A., Raxworthy, C. J., Razafimpahanana, A., Schatz, G. E., Vences, M., … Zjhra, M. L. (2008). Aligning conservation priorities across taxa in Madagascar with high-resolution planning tools. Science, 320(5873), 222–226. https://doi.org/10.1126/science.1155193

Kujala, H., Lahoz-Monfort, J. J., Elith, J., & Moilanen, A. (2018). Not all data are equal: Influence of FCUP 213 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

data type and amount in spatial conservation prioritisation. Methods in Ecology and Evolution, 9(11), 2249–2261. https://doi.org/10.1111/2041-210X.13084

Kukkala, A. S., Arponen, A., Maiorano, L., Moilanen, A., Thuiller, W., Toivonen, T., Zupan, L., Brotons, L., & Cabeza, M. (2016). Matches and mismatches between national and EU-wide priorities: Examining the Natura 2000 network in vertebrate species conservation. Biological Conservation, 198, 193–201. https://doi.org/10.1016/j.biocon.2016.04.016

Kukkala, A. S., & Moilanen, A. (2013). Core concepts of spatial prioritisation in systematic conservation planning. Biological Reviews, 88(2), 443–464. https://doi.org/10.1111/brv.12008

Laitila, J., & Moilanen, A. (2012). Use of many low-level conservation targets reduces high-level conservation performance. Ecological Modelling, 247, 40–47. https://doi.org/10.1016/j.ecolmodel.2012.08.010

Leathwick, J., Moilanen, A., Francis, M., Elith, J., Taylor, P., Julian, K., Hastie, T., & Duffy, C. (2008). Novel methods for the design and evaluation of marine protected areas in offshore waters. Conservation Letters, 1(2), 91–102. https://doi.org/10.1111/j.1755-263x.2008.00012.x

Lehtomäki, J., & Moilanen, A. (2013). Methods and workflow for spatial conservation prioritization using Zonation. Environmental Modelling and Software, 47, 128–137. https://doi.org/10.1016/j.envsoft.2013.05.001

Lehtomäki, J., Tomppo, E., Kuokkanen, P., Hanski, I., & Moilanen, A. (2009). Applying spatial conservation prioritization software and high-resolution GIS data to a national-scale study in forest conservation. Forest Ecology and Management, 258(11), 2439–2449. https://doi.org/10.1016/j.foreco.2009.08.026

Levin, N., Ali, S., Crandall, D., & Kark, S. (2019). World Heritage in danger: Big data and remote sensing can help protect sites in conflict zones. Global Environmental Change, 55(September 2018), 97–104. https://doi.org/10.1016/j.gloenvcha.2019.02.001

Maron, M., Simmonds, J. S., Watson, J. E. M., Sonter, L. J., Bennun, L., Griffiths, V. F., Quétier, F., Hase, A. Von, Edwards, S., Rainey, H., Bull, J. W., Savy, C. E., Victurine, R., Kiesecker, J., Puydarrieux, P., Stevens, T., Cozannet, N., & Jones, J. P. G. (2019). Global no net loss of natural ecosystems. Nature Ecology & Evolution, 4, 46–49.

214 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Martnez-Harms, M. J., & Balvanera, P. (2012). Methods for mapping ecosystem service supply: A review. International Journal of Biodiversity Science, Ecosystem Services and Management, 8(1– 2), 17–25. https://doi.org/10.1080/21513732.2012.663792

Meller, L., Cabeza, M., Pironon, S., Barbet-Massin, M., Maiorano, L., Georges, D., & Thuiller, W. (2014). Ensemble distribution models in conservation prioritization: From consensus predictions to consensus reserve networks. Diversity and Distributions, 20(3), 309–321. https://doi.org/10.1111/ddi.12162

Mengist, W., Soromessa, T., & Legese, G. (2020). Ecosystem services research in mountainous regions: A systematic literature review on current knowledge and research gaps. Science of the Total Environment, 702, 134581. https://doi.org/10.1016/j.scitotenv.2019.134581

Milcu, A. I., Hanspach, J., Abson, D., & Fischer, J. (2013). Cultural ecosystem services: A literature review and prospects for future research. Ecology & Society, 18(3), 44–88. https://doi.org/10.5751/ES-05790-180344

Moilanen, A. (2007). Landscape Zonation, benefit functions and target-based planning: Unifying reserve selection strategies. Biological Conservation, 134(4), 571–579. https://doi.org/10.1016/j.biocon.2006.09.008

Moilanen, A. (2013). Planning impact avoidance and biodiversity offsetting using software for spatial conservation prioritisation. Wildlife Research, 40(2), 153–162. https://doi.org/10.1071/WR12083

Moilanen, A., Anderson, B. J., Arponen, A., Pouzols, F. M., & Thomas, C. D. (2013). Edge artefacts and lost performance in national versus continental conservation priority areas. Diversity and Distributions, 19(2), 171–183. https://doi.org/10.1111/ddi.12000

Moilanen, A., Anderson, B. J., Eigenbrod, F., Heinemeyer, A., Roy, D. B., Gillings, S., Armsworth, P. R., Gaston, K. J., & Thomas, C. D. (2011). Balancing alternative land uses in conservation prioritization. Ecological Applications, 21(5), 1419–1426. https://doi.org/10.1890/10-1865.1

Moilanen, A., & Arponen, A. (2011). Administrative regions in conservation: Balancing local priorities with regional to global preferences in spatial planning. Biological Conservation, 144(5), 1719– 1725. https://doi.org/10.1016/j.biocon.2011.03.007 FCUP 215 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Moilanen, A., Franco, A. M. A. A., Early, R. I., Fox, R., Wintle, B., & Thomas, C. D. (2005). Prioritizing multiple-use landscapes for conservation: Methods for large multi-species planning problems. Proceedings of the Royal Society B: Biological Sciences, 272(1575), 1885–1891. https://doi.org/10.1098/rspb.2005.3164

Moilanen, A., Pouzols, F. M., Meller, L., Veach, V., Arponen, A., Leppänen, J., & Kujala, H. (2014). Zonation - Spatial Conservation Planning Methods and Software Version 4 User Manual. Univ. Helsinki.

Moilanen, A., & Wintle, B. A. (2007). The boundary-quality penalty: A quantitative method for approximating species responses to fragmentation in reserve selection. Conservation Biology, 21(2), 355–364. https://doi.org/10.1111/j.1523-1739.2006.00625.x

Nahuelhual, L., Carmona, A., Lozada, P., Jaramillo, A., & Aguayo, M. (2013). Mapping recreation and ecotourism as a cultural ecosystem service: An application at the local level in Southern Chile. Applied Geography, 40, 71–82. https://doi.org/10.1016/j.apgeog.2012.12.004

Naidoo, R., Balmford, A., Ferraro, P. J., Polasky, S., Ricketts, T. H., & Rouget, M. (2006). Integrating economic costs into conservation planning. Trends in Ecology and Evolution, 21(12), 681–687. https://doi.org/10.1016/j.tree.2006.10.003

Newbold, T., Hudson, L. N., Hill, S. L. L., Contu, S., Lysenko, I., Senior, R. A., Börger, L., Bennett, D. J., Choimes, A., Collen, B., Day, J., De Palma, A., Díaz, S., Echeverria-Londoño, S., Edgar, M. J., Feldman, A., Garon, M., Harrison, M. L. K., Alhusseini, T., … Purvis, A. (2015). Global effects of land use on local terrestrial biodiversity. Nature, 520(7545), 45–50. https://doi.org/10.1038/nature14324

OECD/SWAC. (2014). An Atlas of the Sahara-Sahel: Geography, Economics and Security. OECD Publishing. https://doi.org/http://dx.doi.org/10.1787/9789264222359-en

Ogada, D. L., Keesing, F., & Z. Virani, M. (2012). Dropping dead : causes and consequences of vulture population declines worldwide. Annals Of The New York Academy Of Sciences, 1249, 57–71. https://doi.org/10.1111/j.1749-6632.2011.06293.x

Ospina, G. A. (2006). War and ecotourism in the National Parks of Colombia: some reflections on the public risk and adventure. International Journal of Tourism Research, 8(3), 241–246.

216 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

https://doi.org/10.1002/jtr.572

Paul, C., Hanley, N., Meyer, S. T., Fürst, C., Weisser, W. W., & Knoke, T. (2020). On the functional relationship between biodiversity and economic value. Science Advances, 6(January), eaax7712.

Peña, L., Casado-Arzuaga, I., & Onaindia, M. (2015). Mapping recreation supply and demand using an ecological and a social evaluation approach. Ecosystem Services, 13, 108–118. https://doi.org/10.1016/j.ecoser.2014.12.008

Peres-Neto, P. R., Jackson, D. A., & Somers, K. M. (2003). Giving meaningful interpretation to ordination axes: Assessing loading significance in principal component analysis. Ecology, 84(9), 2347–2363. https://doi.org/10.1890/00-0634

Pouzols, F. M., Toivonen, T., Minin, E. Di, Kukkala, A. S., Kullberg, P., Kuustera, J., Lehtomaki, J., Tenkanen, H., Verburg, P. H., & Moilanen, A. (2014). Global protected area expansion is compromised by projected land-use and parochialism. Nature, 516(7531), 383–386. https://doi.org/10.1038/nature14032

Powers, R. P., & Jetz, W. (2019). Global habitat loss and extinction risk of terrestrial vertebrates under future land-use-change scenarios. Nature Climate Change, 9(4), 323–329. https://doi.org/10.1038/s41558-019-0406-z

Saarinen, J. (2019). What are wilderness areas for? Tourism and political ecologies of wilderness uses and management in the Anthropocene. Journal of Sustainable Tourism, 27(4), 472–487. https://doi.org/10.1080/09669582.2018.1456543

Safriel, U., Adeel, Z., Niemeijer, D., Puigdefabregas, J., White, R., Lal, R., Winslow, M., Ziedler, J., Prince, S., Archer, E., King, C., Shapiro, B., Wessels, K., Nielsen, T., Portnov, B., Reshef, I., Thonell, J., Lachman, E., & Mcnab, D. (2005). Chapter 22: Dryland Systems. Ecosystems and Human Well-Being: Current State and Trends, Volume 1, 625–664.

Santarém, F., Campos, J. C., Pereira, P., Hamidou, D., Saarinen, J., & Brito, J. C. (2018). Using multivariate statistics to assess ecotourism potential of water-bodies: A case-study in Mauritania. Tourism Management, 67, 34–46. https://doi.org/10.1016/j.tourman.2018.01.001

Santarém, F., & Paiva, F. (2015). Conserving desert biodiversity through ecotourism. Tourism FCUP 217 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Management Perspectives, 16, 176–178. https://doi.org/10.1016/j.tmp.2015.07.016

Santarém, F., Pereira, P., Saarinen, J., & Brito, J. C. (2019). New method to identify and map flagship fleets for promoting conservation and ecotourism. Biological Conservation, 229(July 2018), 113– 124. https://doi.org/10.1016/j.biocon.2018.10.017

Santarém, F., Saarinen, J., & Brito, J. C. (2020a). Desert Conservation and Management: Ecotourism. In M. I. Goldstein & D. A. B. T.-E. of the W. B. DellaSala (Eds.), Encyclopedia of the World’s Biomes (pp. 259–273). Elsevier. https://doi.org/https://doi.org/10.1016/B978-0-12-409548- 9.11827-5

Santarém, F., Saarinen, J., & Brito, J. C. (2020b). Mapping and analysing cultural ecosystem services in conflict areas. Ecological Indicators, 110(July 2019), 105943. https://doi.org/10.1016/j.ecolind.2019.105943

Satz, D., Gould, R. K., Chan, K. M. A., Guerry, A., Norton, B., Satterfield, T., Halpern, B. S., Levine, J., Woodside, U., Hannahs, N., Basurto, X., & Klain, S. (2013). The challenges of incorporating cultural ecosystem services into environmental assessment. Ambio, 42(6), 675–684. https://doi.org/10.1007/s13280-013-0386-6

Scholte, S. S. K., Daams, M., Farjon, H., Sijtsma, F. J., van Teeffelen, A. J. A., & Verburg, P. H. (2018). Mapping recreation as an ecosystem service: Considering scale, interregional differences and the influence of physical attributes. Landscape and Urban Planning, 175(March), 149–160. https://doi.org/10.1016/j.landurbplan.2018.03.011

Sharp, R., Chaplin-Kramer, R., Wood, S., Guerry, A., Tallis, H., Ricketts, T., Nelson, E., Ennaanay, D., Wolny, S., Olwero, N., Vigerstol, K., Pennington, D., Mendoza, G., Aukema, J., Foster, J., Forrest, J., Cameron, D., Arkema, K., Lonsdorf, E., … Bierbower, W. (2014). InVEST User Guide. National Capital Project. http://data.naturalcapitalandresilienceplatform.org/invest- releases/documentation/current_release/index.html

Small, N., Munday, M., & Durance, I. (2017). The challenge of valuing ecosystem services that have no material benefits. Global Environmental Change, 44, 57–67. https://doi.org/10.1016/j.gloenvcha.2017.03.005

Strand, J., Soares-filho, B., Costa, M. H., Oliveira, U., Ribeiro, S. C., Pires, G. F., Oliveira, A., Rajão,

218 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

R., May, P., Hoff, R. Van Der, Siikamäki, J., Seroa, R., & Toman, M. (2018). Spatially explicit valuation of the Brazilian Amazon Forest ’ s Ecosystem Services. Nature Sustainability, 1(November), 657–664. https://doi.org/10.1038/s41893-018-0175-0

Taylor, N. T., Davis, K. M., Abad, H., McClung, M. R., & Moran, M. D. (2017). Ecosystem services of the Big Bend region of the Chihuahuan Desert. Ecosystem Services, 27, 48–57. https://doi.org/10.1016/j.ecoser.2017.07.017

Teff‐Seker, Y., & Orenstein, D. E. (2019). The ‘desert experience’: Evaluating the cultural ecosystem services of drylands through walking and focusing. People and Nature, 1(1), 234–248. https://doi.org/10.1002/pan3.28

Teoh, S. H. S., Symes, W. S., Sun, H., Pienkowski, T., & Carrasco, L. R. (2019). A global meta-analysis of the economic values of provisioning and cultural ecosystem services. Science of the Total Environment, 649, 1293–1298. https://doi.org/10.1016/j.scitotenv.2018.08.422

Tilman, D., Clark, M., Williams, D. R., Kimmel, K., Polasky, S., & Packer, C. (2017). Future threats to biodiversity and pathways to their prevention. Nature, 546(7656), 73–81. https://doi.org/10.1038/nature22900

Toivonen, T., Heikinheimo, V., Fink, C., Hausmann, A., Hiippala, T., Järv, O., Tenkanen, H., & Di Minin, E. (2019). Social media data for conservation science: A methodological overview. Biological Conservation, 233(January), 298–315. https://doi.org/10.1016/j.biocon.2019.01.023

UNDP. (2019). Human Development Report 2019: Beyond Income, Beyond Averages, Beyond Today - Inequalities in Human Development in the 21st Century (Human Development Report). United Nations Development Programme. https://www.un-ilibrary.org/economic-and-social- development/human-development-report-2019_838f78fd-en

UNEP. (2006a). Global Deserts Outlook (E. Ezcurra (ed.)). United Nations Environment Programme. http://www.unep.org/geo/GDOutlook/

UNEP. (2006b). Tourism and Deserts: A Practical Guide to Managing the Social and Environmental Impacts in the Desert Recreation Sector. United Nations Environment Programme.

UNESCO. (2003a). The Sahara of Cultures and People: Towards a strategy for the sustainable FCUP 219 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

development of tourism in the Sahara, in the context of combating poverty (Vol. 3, Issue 12). https://doi.org/10.1017/S0003598X00003768

UNESCO. (2003b). Towards a strategy for the sustainable development of tourism in the Sahara, in the context of combating poverty. In The Sahara of Cultures and People (Issue July).

United Nations. (2019). United Nations Mission for the Referendum in Western Sahara. https://minurso.unmissions.org/

UNWTO. (2019). Yearbook of Tourism Statistics, Data 2013-2017, 2019 Edition. https://www.e- unwto.org/doi/book/10.18111/9789284420414

Vale, C. G., Pimm, S. L., & Brito, J. C. (2015). Overlooked mountain rock pools in deserts are critical local hotspots of biodiversity. PLoS ONE, 10(2), e0118367. https://doi.org/10.1371/journal.pone.0118367

Venter, O., Sanderson, E.W., Magrach, A., Allan, J.R., Beher, J., Jones, K.R., Possingham, H.P., Laurance, W.F., Wood, P., Fekete, B.M., Levy, M.A., Watson, J.E., 2018. Last of the Wild Project, Version 3 (LWP-3): 2009 Human Footprint, 2018 Release.

Venter, O., Sanderson, E.W., Magrach, A., Allan, J.R., Beher, J., Jones, K.R., Possingham, H.P., Laurance, W.F., Wood, P., Fekete, B.M., Levy, M.A., Watson, J.E.M., 2016a. Global terrestrial Human Footprint maps for 1993 and 2009. Sci. Data 3, 1–10. https://doi.org/10.1038/sdata.2016.67a

Villa, F., Bagstad, K. J., Voigt, B., Johnson, G. W., Portela, R., Honzák, M., & Batker, D. (2014). A methodology for adaptable and robust ecosystem services assessment. PLoS ONE, 9(3). https://doi.org/10.1371/journal.pone.0091001

Waldron, A., Miller, D. C., Redding, D., Mooers, A., Kuhn, T. S., Nibbelink, N., Roberts, J. T., Tobias, J. A., & Gittleman, J. L. (2017). Reductions in global biodiversity loss predicted from conservation spending. Nature, 551(7680), 364–367. https://doi.org/10.1038/nature24295

Waldron, A., Mooers, A. O., Miller, D. C., Nibbelink, N., Redding, D., & Kuhn, T. S. (2013). Targeting global conservation funding to limit immediate biodiversity declines. Proceedings of the National Academy of Sciences USA, 110(29), 1–5. https://doi.org/10.5061/dryad.p69t1

220 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Walther, O. J. (2017). Wars and Conflicts in the Sahara-Sahel. In West African Papers (10, Issue 10). https://www.oecd-ilibrary.org/content/paper/8bbc5813-en

Watson, J. E. M., Venter, O., Lee, J., Jones, K. R., Robinson, J. G., Possingham, H. P., & Allan, J. R. (2018). Protect the last of the wild. Nature, 563, 27–30. https://doi.org/10.1038/ d41586-018- 07183-6

Whitehead, A. L., Kujala, H., Ives, C. D., Gordon, A., Lentini, P. E., Wintle, B. A., Nicholson, E., & Raymond, C. M. (2014). Integrating biological and social values when prioritizing places for biodiversity conservation. Conservation Biology, 28(4), 992–1003. https://doi.org/10.1111/cobi.12257

Willemen, L., Barger, N. N., Brink, B., Cantele, M., Erasmus, B. F. N., Fisher, J. L., Gardner, T., Holland, T. G., Kohler, F., Kotiaho, J. S., Maltitz, G. P. Von, Nangendo, G., Pandit, R., Parrotta, J. A., Potts, M. D., Prince, S. D., Sankaran, M., Brainich, A., Montanarella, L., & Scholes, R. (2020). How to halt the global decline of lands. Nature Sustainability, 3, 164–166. http://dx.doi.org/10.1038/s41893-020-0477-x

Winkler, D. E., & Brooks, E. (2020). Tracing Extremes across Iconic Desert Landscapes: Socio- Ecological and Cultural Responses to Climate Change, Water Scarcity, and Wildflower Superblooms. Human Ecology, 48(2), 211–223. https://doi.org/10.1007/s10745-020-00145-5

Wintle, B. A., Kujala, H., Whitehead, A., Cameron, A., Veloz, S., Kukkala, A., Moilanen, A., Gordon, A., Lentini, P. E., Cadenhead, N. C. R., & Bekessy, S. A. (2019). Global synthesis of conservation studies reveals the importance of small habitat patches for biodiversity. Proceedings of the National Academy of Sciences of the United States of America, 116(3), 909–914. https://doi.org/10.1073/pnas.1813051115

Wolff, S., Schulp, C. J. E., & Verburg, P. H. (2015). Mapping ecosystem services demand: A review of current research and future perspectives. Ecological Indicators, 55, 159–171. https://doi.org/10.1016/j.ecolind.2015.03.016

Wood, S. L. R., Jones, S. K., Johnson, J. A., Brauman, K. A., Chaplin-Kramer, R., Fremier, A., Girvetz, E., Gordon, L. J., Kappel, C. V., Mandle, L., Mulligan, M., O’Farrell, P., Smith, W. K., Willemen, L., Zhang, W., & DeClerck, F. A. (2018). Distilling the role of ecosystem services in the Sustainable Development Goals. Ecosystem Services, 29(December), 70–82. FCUP 221 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

https://doi.org/10.1016/j.ecoser.2017.10.010

World Tourism Organization. (2014). Towards Measuring the Economic Value of Wildlife Watching Tourism in Africa – Briefing Paper. In Towards Measuring the Economic Value of Wildlife Watching Tourism in Africa – Briefing Paper.

Yakum, M. N., Ateudjieu, J., Guenou, E., Walter, E. A., Ram, M., Debes, A. K., Njimbia, A. C., Nafack, S. S., & Sack, D. A. (2017). Health seeking behaviour among suspected cases of cholera in Cameroonian health districts in Lake Chad basin. BMC Research Notes, 10(1), 1–7. https://doi.org/10.1186/s13104-017-2756-9

Zulian, G., Polce, C., & Maes, J. (2014). ESTIMAP: A GIS-based model to map ecosystem services in the European Union. Annali Di Botanica, 4(March), 1–7. https://doi.org/10.4462/annbotrm-11807

222 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Article VI: Vulnerability of desert tourism to climate change: lessons from the Sahara-Sahel

Frederico Santarém1,2, Duarte V. Gonçalves 3, Jarkko Saarinen4,5, José Carlos Brito1,2

1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 - CIIMAR, Interdisciplinary Centre of Marine and Environmental Research, University of Porto, Terminal de Cruzeiros do Porto de Leixões, Avenida General Norton de Matos, S/N, 4450-208 Matosinhos, Portugal 4 - Geography Research Unit, University of Oulu, Finland 5 - School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa FCUP 223 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Abstract

The Sahara-Sahel is the largest warm region in the world with sustainable tourism potential. There, however, climate change impacts are predicted to be strong, yet they remain understudied. Using three bioclimatic variables related to temperature and precipitation and eight global climate models, we investigated the impacts of climate change on Sahara-Sahel tourism hotspots under two socioeconomic paths for development. Results show that precipitation shifts will be strong in south-easternmost areas, where specialized desert-tourism will be challenged. Numerous vulnerable tourism hotspots will be severely impacted by extreme temperatures, particularly under the worst-case climate change scenario. Some countries – like Burkina Faso and Mali– will have their tourism hotspots severely impacted, with temperatures challenging human survival all the yearlong. New development and sustainable management policies will be needed to adapt to extreme climate shifts. The framework is replicable to other global regions where climate change is forecasted to impact tourism.

Keywords Desert tourism; Sahara-Sahel; global warming; Africa; sustainable development; vulnerability

224 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Introduction

Tourism has been growing at a fast pace (UNWTO, 2020), contributing USD8.9 trillion to the global economy (10.3% of the 2019 global GDP; WTTC, 2020) before the COVID-19 pandemic. The travel and tourism sectors accounted for 330 million jobs in 2019, representing about 10% of the global employment (WTTC, 2020). Many countries rely heavily on tourism, particularly in Africa where it generated USD 165.8 billion and 24.6 million employment in 2019 (WTTC, 2020). As such, the United Nations have been positioning tourism as an important contributor to the Sustainable Development Goals (SDGs), mainly to overcome poverty and inequalities, while conserving biological resources (Hall, 2019; UNWTO, 2015). The potential of tourism to contribute positively to the SDGs and global biodiversity conservation can be expected to continue when global tourism resumes after the COVID crisis (Gössling, et al., 2020; Rogerson & Baum, 2020; Saarinen, 2021). Global climate change poses increasing risks to the tourism economy (Bosello et al., 2006; Lamperti et al., 2019), human livelihoods (Bosello et al., 2006; Mora et al., 2018), and entire ecosystems on which everything depends on (Mooney et al., 2009). Among 89 analysed parameters of human health, water, food, economy, security, and infrastructures, it was found that tourism is directly impacted by 10 different types of climate change impacts (Mora et al., 2018). In this respect, climate change impacts on tourism can be considered to be profound and far-reaching and will challenge sustainable development for the years to come (Becken, 2013, 2013; Gössling & Scott, 2018; Kaján & Saarinen, 2013; Scott et al., 2012, 2019; Scott, Hall, et al., 2016). In the Sahara-Sahel, the key vulnerabilities of the sector to climate change relate to: destinations resilience, water competition in water-scarce regions, increased energy and food supply costs, tourist demands at destinations, increased social inequality and unrest, and tourism competitiveness for less good spots available for visitation (see Scott et al., 2019; Tervo-Kankare, Kaján & Saarinen, 2018). Furthermore, climate change will increasingly affect travel patterns in the future (Fisichelli et al., 2015), with strong implications forecasted for all the tourism industry sub-sectors (Gössling & Higham, 2020). These issues impact travel and tourism activities at a large extent that still need to be properly quantified. Besides irreversible ecosystem impacts and massive human displacement, the likely future increase of 2ºC will also render many regions too hot to visit safely or comfortably (at least seasonally; IPCC, 2018). Visitor comfort starts to cease above 30ºC under well-ventilated outdoor conditions (Maddison, 2001), even in hot countries (Gössling & Hall, 2006). The human body is FCUP 225 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel vulnerable to heat stroke at higher temperatures (e.g. >45ºC) or if exposed to the sun for too long without proper hydration (Sherwood & Huber, 2010). Still, temperature-related health risks remain unnoticed in the tourism-climate nexus literature. Also, precipitation is projected to decrease substantially in subtropical dry regions, intensifying competition for an already scarce resource (Cherlet et al., 2018; IPCC, 2018; Ogunrinde et al., 2020). Other regions should experience increases in precipitation and/or extreme rainfall and floods, with their own logistic and infrastructural challenges (Becken et al., 2015; Gössling et al., 2006; Hoogendoorn & Fitchett, 2018). Overall, tourism is highly dependent on environmental quality and climate change may have serious impacts on the attractiveness of certain regions. Thus, while the adaptive capacity of the tourism industry is globally high, this is not the case for destinations where climate change is projected to severely degrade ecosystems, such as drylands (WTTC, 2020). The literature over climate change is increasing, but the consequences of global warming and precipitation shifts on ecosystems and human societies are scarcely investigated (Scott, Hall, et al., 2016), especially in the Sahara-Sahel region with long tradition of tourism. In general, the literature on tourism vulnerability to climate change is extensive and growing at fast pace (Amelung et al., 2007; Becken, 2013; Becken et al., 2020; Dubois & Ceron, 2006; Fang et al., 2017; Gössling et al., 2012; Gössling & Scott, 2018; Hall, 2019; Hewer & Gough, 2018; Kaján & Saarinen, 2013; Scott, 2011; Scott et al., 2012, 2019; Scott, Hall, et al., 2016; Scott & Becken, 2010). Still, there is limited published research on the context of low-income countries and especially from the Sahara-Sahel region (Becken, 2013; Kaján & Saarinen, 2013; Rogerson, 2016; Scott et al., 2019; Scott, Hall, et al., 2016; Scott & Becken, 2010). Indeed, the geographic focus of research on climate change impacts on tourism has been mostly towards the global north (Hoogendoorn & Fitchett, 2018; Rogerson, 2016; Saarinen et al., 2012; Tervo-Kankare et al., 2017). This bias towards the Global North is surprising as the impacts of climate change are anticipated to be very salient in Africa where sustainable tourism is one of the most prominent economic activities (Hoogendoorn & Fitchett, 2018; Scott, Hall, et al., 2016; UNCTAD, 2017). For instance, Egypt and Nigeria are amongst the largest contributors to Africa GDP, and Tunisia is the fastest growing tourism country in the continent (Tunisia’ Travel & Tourism GDP expanded 12.9% in 2019; WTTC, 2020). Therefore, it is urgent to study the vulnerability of African tourism to climate change. In relation to climate change, the Sahara-Sahel may be under the most vulnerable situation, as the region is subjected to many armed conflicts that threaten biodiversity and people alike (Brito et al., 2018; Daskin & Pringle, 2018; Dioko & Harrill, 2019; Walther, 2017) and to unsustainable exploitation of resources that severely deplete natural ecosystems (Brito et al.,

226 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

2014, 2018; OECD/SWAC, 2014; Santarém et al., 2020b). These pressures are predicted to increase under climate change (Brito et al., 2014; Schleussner et al., 2016). And while a large proportion of its biodiversity is projected to be up-listed in conservation status directly and indirectly because of climate change impacts (Durant et al., 2014; Powers & Jetz, 2019), the region remains among the least studied (Brito et al., 2014; Durant et al., 2012, 2014) and the most underfunded for conservation (Waldron et al., 2013). Sahara-Sahel is among the last global wild places (Di Marco et al., 2019; Watson et al., 2018) partially managed by indigenous people (Garnett et al., 2018), offering chances to tourists seeking ‘last-chance-to-see’ wild places (Saarinen, 2019). Cultural ecosystem services hotspots that benefit people from recreation and tourism experiences have been identified in the region (Santarém et al., 2020b) but they remain largely unprotected (Brito et al., 2016). Most of the Sahara-Sahel countries are among the world’s poorest (UNDP, 2019) and least visited (UNWTO, 2020), fostering a continuing regional ecosystem crisis that fail to cease and will be exacerbated by global warming (Brito et al., 2014, 2018; Schleussner et al., 2016). Desert-tourism is one of the most promising economic activities for biological conservation and sustainable development in the region (Santarém et al., 2020a; Santarém & Paiva, 2015; UNEP, 2006b; UNESCO, 2003; Weaver, 2001) but is vulnerable to climate change. Thus, geographical-explicit assessments of the impact of climate change in the Sahara-Sahel is urgently needed. A few studies have used the Tourism Climate Index for assessing tourism comfort (e.g. Mushawemhuka, Fitchett, and Hoogendoorn 2020; Amelung, Nicholls, and Viner 2007; Amelung and Nicholls 2014; Hein, Metzger, and Moreno 2009; Fitchett, Robinson, and Hoogendoorn 2017), despite the critiques to the accuracy and reliability of the methodology (Scott, Rutty, et al., 2016). Still, no study has used the most refined bioclimatic data (Fick & Hijmans, 2017) to quantitatively assess and map climate change impacts on tourism assets (Becken et al., 2015). Overcoming this issue is of paramount interest to regional policymakers and tourism developers and managers who need to cope with new challenges for tourism and other cultural ecosystem services (Monz et al., 2020). Here we perform a spatially explicit assessment of climate change impacts over tourism hotspots in the largest warm arid region in the world: the Sahara-Sahel in northern Africa where tourism has played a very prominent economic and social role in the past. We used refined bioclimatic data and updated global climate models to understand where impacts on tourism hotspots will be more severe. In particular, we wanted to answer two questions: 1) how will climate change affect weather tolerability for tourism in Sahara-Sahel and challenge human survival? and 2) how will countries be affected by climate change impacts relative to each other? Answering these questions will help to design and direct adequate adaptation policies to overcome ongoing FCUP 227 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel and future climatic shifts affecting tourism hotspots and human livelihoods in the region. Regional and international policymakers, planners and businesses may find here a framework to better prepare for one of the most widespread challenges the Humanity faces – climate change – and its estimated impacts for tourism-dependent socioeconomic development in future.

Methods

Study Area The Sahara-Sahel is the largest warm region in the world (≈11,200,000 km2) and includes the African ecoregions of the Sahara Desert (≈8,200,000km2) and its contiguous Sahel arid area (≈3,000,000km2) (Fig. 1; Dinerstein et al., 2017). We have divided the study area into 4120 grid cells of 0.5-degree resolution (WGS84 coordinate reference system). The region underwent cyclic climatic fluctuations and geological events that shaped climacteric conditions in the region in ancient timescales (Couvreur et al., 2020). The climate is heterogeneous, with the average annual temperature ranging from 9.6ºC to 30.6ºC and the average annual total precipitation from 0 to 1184 mm (Fig.1). The region is rich in natural and cultural heritage that benefit local communities and international tourists (Hosni, 2000; Santarém et al., 2020a, 2020b; UNEP, 2006a; UNESCO, 2003), but it is vulnerable to unsustainable development and climate change (Brito et al., 2014, 2018). Climate has been warming at an alarming rate in the Sahara, and the Sahel has been suffering dramatic shifts responsible for decadal severe storms and long-lasting droughts (Cook & Vizy, 2015; Dong & Sutton, 2015; Taylor et al., 2017). The massive drought and famine in the 70s is a stark reminder of how quickly and severely conditions for human survivability can decline in the region (Ogunrinde et al., 2020)

228 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 4.7. Location of the Sahara-Sahel within North Africa and its countries (top), and the present climate conditions in the study area, depicted by selected bioclimatic variables (bottom). BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic variables were obtained at 10-minute resolution from the WorldClim version 2.1 climate data for the years 1970-2000 (Fick & Hijmans, 2017) at https://www.worldclim.org. Country codes: ALG: Algeria; BUF: Burkina Faso; CAM: Cameroon; CHA: Chad; EGY: Egypt; ERI: Eritrea; ETH: Ethiopia; LIB: Libya; MAL: Mali; Mauritania: MAU; MOR: Morocco; NIG: Niger; NIA: Nigeria; SEN: Senegal; SSU: South Sudan; SUD: Sudan; and TUN: Tunisia.

Climate Change Models To understand the implications of future climate shifts on tourism hotspots, we collected historical and future (years 2081-2100) data of bioclimatic variables related to annual mean temperature (BIO 01), maximum temperature of warmest month (BIO 05), and annual precipitation (BIO 12) from WorldClim database at 10-min spatial resolution (https://www.worldclim.org/; Fick & Hijmans, 2017). While these may not represent all dimensions and detail of climate change, they are adequate surrogates, considering we are concerned with broad-scale and multi-decade trends. Future climate data (BIO 01, 05 and 12) were collected from eight different Global Climate Models (GCMs; Eyring et al., 2016). Two Shared Socioeconomic Pathways (SSPs) with two different IPCC Representative Concentration Pathways (RCPs) (SSP1-2.6 and SSP5-8.5; Riahi et al., 2017) defined under the 6th Coupled Model Intercomparison Project (CMIP6; O’Neill et al., FCUP 229 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

2016, Eyring et al., 2016; Gidden et al., 2019) were considered (Table 1). SSPs represent distinct pathways that examine how global society, demographics and economics might change over the next century and how the climate goals of the Paris Agreement could be met (Riahi et al., 2017). They are based on different narratives that describe alternative socioeconomic developments for the future. SSP1 represents a world of sustainability-focused growth and equality with low- resource and low-energy consumption (van Vuuren et al., 2017). On the opposite side SSP5 represents a world of rapid and unconstrained growth in economic output and energy use (Kriegler et al., 2017). RCPs represent different emissions, concentration and radiative forcing projections leading to a large range of global warming levels. RCP2.6 is on the low end of the climate change scenarios in terms of emissions and radiative forcing within SSP1 (van Vuuren et al., 2011), whereas RCP8.5 corresponds to the pathway with the highest greenhouse gas emissions within SSP5 (Riahi et al., 2011). The two efforts were designed to be complementary: the RCPs define pathways for greenhouse gas concentrations and the amount of warming that could occur by the end of the 21st century, whereas the SSPs determine which reductions in emissions will or will not be achieved. Specifically, each SSP looks at how each different RCP could be achieved within the context of the underlying socioeconomic characteristics and shared policy assumptions (Kriegler et al., 2014). In essence, for our work SSP1 RCP 2.6 is the most optimistic scenario, whereas SSP5 RCP 8.5 is the most pessimistic. For simplicity, hereafter we will follow the WorldClim in calling these scenarios and pathways as SSP126 and SSP585. For a thorough explanation of SSPs, see Kriegler et al. (2017), Riahi et al. (2017), and van Vuuren et al. (2017).

Assembling models for the future Future climate predictions for Sahara-Sahel (Fig. S1) were obtained by averaging the predictions of the eight high-resolution GCMs, using the Average (mean) function in ArcMap 10.1 (ESRI, 2012). We then upscaled the averaged values from 10-minute to 0.5-degree resolution (WGS84 coordinate reference system).

230 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table 4.5. Global Climate Models (GCMs) for two Shared Socioeconomic Pathways (SSPs 126 and 585; Riahi et al., 2017) for the years 2081-2100 used in this study and the institutions responsible for developing those GCMs. The codes, model full names (when available), institutions responsible for the models and the main reference are provided. Models and institutional data retrieved from the Coupled Model Intercomparison Project 6 (CMIP6; Eyring et al., 2016) of the World Climate Research Programme (WCRP)’s Working Group of Coupled Modelling (WGCM) at https://www.wcrp-climate.org/wgcm-cmip. For full data sources consult the World Data Center for Climate (WDCC) at https://cera.-www.dkrz.de. Bioclimatic variables for all GCMs were downloaded at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org. Code Model Full Institution Reference Name BCC-CSM2- Beijing Climate BCC: Beijing Climate Center, China Wu et al. (2019) MR Center Climate System Model CNRM-CM6- CNRM-CERFACS: Centre National Voldoire et al. 1 de Recherches Meteorologiques, (2019) France; Centre Européen de Recherche et de Formation Avancée, France CNRM- CNRM-CERFACS: Centre National Séférian et al., ESM2-1 de Recherches Meteorologiques, (2019) France; Centre Européen de Recherche et de Formation Avancée CanESM5 Canadian Earth CCCma: Canadian Centre for Swart et al. (2019) System Model Climate Modelling and Analysis, Version 5 Environment and Climate Change, Canada IPSL-CM6A- IPSL: Institut Pierre Simon Laplace, Boucher et al. LR Paris (2020) MIROC- Model for MIROC: JAMSTEC (Japan Agency Hajima et al. (2020) ES2L Interdisciplinary for Marine-Earth Science and Research on Technology), AORI (Atmosphere and Climate, Earth Ocean Research Institute, The System version University of Tokyo), NIES (National 2 for Long-term Institute for Environmental Studies), simulations and R-CCS (RIKEN Center for Computational Science), Japan MIROC6 Model for MIROC: JAMSTEC (Japan Agency Tatebe et al. (2018) Interdisciplinary for Marine-Earth Science and Research on Technology), AORI (Atmosphere and Climate version Ocean Research Institute, The 6 University of Tokyo), NIES (National Institute for Environmental Studies), and R-CCS (RIKEN Center for Computational Science), Japan MRI-ESM2-0 Meteorological MRI: Meteorological Research Yukimoto et al. Research Institute (2019) Institute Earth System Model version 2.0

FCUP 231 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Future climate differences We assessed future climatic shifts in the Sahara-Sahel by accounting for future differences in temperatures and precipitation from present conditions. We subtracted present values of temperatures (BIO 01 and BIO 05) and precipitation (BIO12) from future mean values to obtain averaged differences between present and future projected climate. Because projected climatic shifts can vary widely and consequences can be much larger than previously predicted, we also accounted for uncertainties in maximum differences between present and future mean projected climate (Fig. S2). Maximum differences were calculated as future averages plus the standard deviation minus present climatic conditions for the three bioclimatic variables.

Endangered tourism hotspots We evaluated the proportion of tourism hotspots in each country that is or will be impacted by changes in temperature and precipitation levels. To do so, we counted the number of 0.5- degree pixels of tourism hotspots falling under each of the following thresholds: >30ºC for annual mean temperature (BIO 01); >45ºC for maximum temperature of warmest month (BIO 05); and <200mm and >1000mm for annual precipitation (BIO 12). These levels define the threshold of survival for a fit human (Sherwood & Huber, 2010) and a threshold near the heat stroke risk (Bynum et al., 1978), as well as the ”desert” precipitation threshold (<200mm) and shifts to wetter- than-Sahel conditions in others (>1000mm; Cherlet et al., 2018). We also explore the trend in tourism hotspots going beyond both temperature thresholds, i.e., we calculated the proportion of tourism hotspots that simultaneously are or will be subject to mean annual temperatures above 30ºC and maximum temperatures of the hottest month above 45ºC under the two SSPs.

Results

Climatic shifts in Sahara-Sahel Future average annual temperatures may reach 32.5ºC under the best-case scenario (SSP126) or 36.7ºC under the worst-case scenario (SSP585; Fig. S1). The maximum temperature of the warmest month may reach 50.4ºC under the best-case scenario and 55.4ºC under the worst-case scenario. Precipitation may reach 1318mm under the best-case scenario and 1649mm under the worst-case scenario. Overall, most of the region will face moderate (1-2ºC differences) to strong (10-11ºC differences) temperature shifts under the SSP126 and SSP585, respectively (Fig.2). Areas close

232 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

to the Atlantic, Indian, and Mediterranean coasts, and to the south limits of Sahel will be less impacted by temperature increases than central areas. Regarding precipitation, most of the Mediterranean and Atlantic coasts will experience precipitation drop up to 100mm below present levels, making them even drier, whereas some southeast areas will experience a dramatic increase of up to 900mm in precipitation. Uncertainties (“Maximum differences” in Fig. 4.7) tend to be more pronounced for bioclimatic data related to temperature (BIO 01 and BIO 05), although they also vary for precipitation (BIO 12), as decreases in precipitation tend to be ameliorated under uncertainty scenarios but increases are more pronounced (Fig. 4.7; Fig. S1-S2). As expected, future climatic differences are more pronounced under the SSP 585 than under the SSP126. Algeria, Mali and some areas of Mauritania and Niger will face the strongest increases in the mean annual temperature (BIO 01) under the worst-case scenario (SSP585). Algeria, Tunisia, Libya, Egypt, and part of north Sudan will face the strongest impacts in temperatures of the warmest month (BIO 05) under the SSP585. About half of Mauritania, Algeria, Libya, and Egypt, and most of Morocco, Senegal and Tunisia will experience the most dramatic drop in precipitation under the SSP585 and southern Chad, Central African Republic, Eritrea, Ethiopia, and parts of eastern Sudan will experience the strongest increases in precipitation under the SSP585 (Fig. 2).

Tourism hotspots threatened by future climate change Currently, very few tourism hotspots are threatened by extreme temperatures during the year (BIO 01) and none during the warmest month of the year (BIO 05; Fig. 3). However, many of them are highly vulnerable and will become threatened in the future by extremes in the average annual temperature and, to some extent, by extreme temperatures during the warmest month of the year. Sahara-Sahel is mostly arid, and it will stay as such in the future (precipitation is less than 200mm for most of the region in the present and in the future), but extreme precipitation shifts are expected to occur in the southeast part of Sahara-Sahel, altering the landscape as a whole and affecting the adequacy for desert-specialized tourists. As expected, changes will be more pronounced under the SSP 585. FCUP 233 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 4.8. Averaged and maximum differences between present and future projected climate. Future climate was averaged over eight Global Climate Models – GCMs, for two Shared Socioeconomic Pathways - SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017). Maximum differences were calculated as future averages plus the standard deviation for the eight GCMs for two SSP minus present climatic conditions. Bioclimatic variables are coded as: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic variables were obtained at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org.

234 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 4.9. Differences between present and future temperature and precipitation levels and their potential impacts on tourism hotspots in Sahara-Sahel. For ease of interpretation, only two thresholds for temperature and four thresholds for precipitation are depicted. Temperatures pushing the limits for visitation are depicted in light red. For precipitation, dark blue depicts changes that may significantly change current ecoregions and, thus, demand for tourists specialised in desert heritage; and light red depicts increased aridity levels. Bioclimatic data for the future were obtained by averaging eight global climate models (GCMs; see Table 4.5) for two Shared Socioeconomic Pathways (SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017). Bioclimatic variables are coded as: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic data for present and the future were obtained at 10-minute resolution from the WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org.

Proportion of endangered tourism hotspots by country Currently, only Mauritania and Mali have some of their tourism hotspots above annual temperatures challenging comfort for tourism activities (6.9% and 2.7% of the total tourism hotspots, respectively), and only Algeria has tourism hotspots threatened by extreme temperatures in the warmest month of the year (0.9% of the total tourism hotspots) (Fig. 4.8 and 4.9; Table S1). In the future, climate models point out to an aggravated situation, where some countries will experience a large proportion of their tourism hotspots severely impacted by climatic shifts. Results for the mean annual temperatures under the best-case scenario (SSP 126) show that all tourism hotspots of Burkina-Faso and many of Mali are threatened, while under the worst- case scenario (SSP 585) the most vulnerable hotspots are located in Burkina Faso, Cameroon, Mali, Mauritania, Niger, Nigeria, Senegal, South Sudan, and Sudan. For the maximum FCUP 235 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel temperatures of the warmest month hotspots in Algeria and Mauritania are threatened under the best-case scenario, and Algeria, Burkina Faso, Mali, Nigeria, and Niger are endangered under the worst-case scenario. Current precipitation levels are not impacting desert attractions in any country while the results for future precipitation under the two analysed SSPs are mixed (Fig. 4; Table S1). Several regions will suffer a decrease in aridity, which will surely impact local human communities and complicate logistics for tourism stakeholders. For the sake of a desert-tourism perspective, we highlight the ones above 1000mm of precipitation, too wet to maintain Sahelian landscapes, and thus losing attractiveness in terms of desert and arid attractions. Under the best-case scenario the issue is marginal but under the worst-case scenario the most affected tourism hotspots will be located in Cameroon, Chad, Ethiopia, and South Sudan.

236 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

PRESENT SSP 126 SSP 585

1

0 BIO

BIO 05 BIO

BIO 12 BIO

Fig. 4.10. Proportion of tourism hotspots in each country that is (Present) or will be impacted (Shared Socioeconomic Pathways: SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017) by high temperature (>30ºC in BIO 01; >45ºC in BIO 05) and will experience dramatic precipitation shifts (>1000mm in BIO 12). For temperature data, light red depicts where climate will impact tourism hotspots the most. Dark blue depicts precipitation levels high enough to alter ecoregions and landscapes, and impact tourist demand for desert attractions. Lesser impacts are depicted in white (temperature and precipitation data) and light blue bars (precipitation only). Bioclimatic variables are coded as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic variables were obtained at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org. See country codes in Fig. 4.7 and the number of tourism hotspots to be impacted under distinct climate levels in Table E1. FCUP 237 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Countries facing the greatest challenges for their tourism hotspots Currently, no country faces large challenges when analysing the proportion of tourism hotspots that have simultaneously annual mean temperatures above 30ºC and maximum temperatures of the warmest month above 45ºC (Fig. 5). Still, the situation will likely change in the future even under the best-case scenario, with Burkina Faso, Mauritania, and Mali becoming motive of concern. The severest conditions are predicted for Burkina Faso, Mali, Niger, and Nigeria under the worst-case scenario, with almost all the tourism hotspots severely impacted by annual mean temperatures and extreme temperatures during the warmest month. Algeria, Chad, Cameroon, Mauritania, and Sudan are also vulnerable under SSP 585, with all of them having more than half of their tourism hotspots threatened by both annual mean temperature above 30ºC and maximum temperature of warmest month above 45ºC.

Fig. 4.11. Proportion of tourism hotspots that are currently (Present) or will be (SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017) subject to mean annual temperatures above 30ºC (BIO 01) and maximum temperatures of the hottest month above 45ºC (BIO 05), threatening their visitation. Bioclimatic variables were obtained at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org. See country codes in Fig. 4.7.

238 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Discussion

Advances to the tourism-climate change nexus After calls to develop maps that quantitatively assess how climate change is impacting tourism (Becken et al., 2015), we have made use of refined bioclimatic data and GCMs projections to understand where tourism management will be most challenged by global warming. Mapping is useful for tourism planning and can highlight future challenges for tourism managers and policymakers to assess vulnerabilities and adapt to a changing climate. Additionally, this is one of the first quantitative spatial studies dedicated to climate change on desert tourism attractions, contributing significant knowledge to this neglected system (Durant et al., 2012). .

Limitations and future improvements The results obtained depended on the chosen bioclimatic data, GCMs, SSPs, RCPs and the upscaling method. We used three bioclimatic variables that are important for desert contexts and adequate surrogates for the climate change process, given the broad spatial scale at which the subject is here addressed. Eight GCMs were used in this work, as they represent the most updated models (Eyring et al., 2016; Gidden et al., 2019; O’Neill et al., 2016) currently available for download in the broadly used WorldClim database (www.worldclim.com). Two SSPs were chosen as they represent opposite poles of predicted socioeconomic developments to explore different possible futures in the context of fundamental future uncertainties (Riahi et al., 2017). There are some limitations in the models used, but these are currently considered the best-in- class for our purposes (see the list of references in Table 1), and their use was here considered as an attempt to deal with uncertainties in climate projections. The choice of parameters in similar future studies should likewise be adequate to the questions and geographical scale at hand. For instance, sea-level rise data may be highly informative to predict impacts on African coastal protected areas (Brito & Naia, 2020) where tourism is a strong driver of economic growth (Brooks et al., 2020). Although frequency and intensity of extreme weather events - such as heat waves and floods - are hard to predict, they are highly relevant for tourism planning and thus can be considered once models become more accurate in predicting their frequency and seasonal timing (Amelung & Nicholls, 2014). Lastly, there are more stochastic events, also potentiated by climate change, that can have a large influence on travelling, such as terrorism events, natural disasters, and pandemics (Gössling et FCUP 239 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel al., 2020; Gössling & Hall, 2006; Gössling & Higham, 2020; Hall et al., 2020; Rosselló et al., 2020). However, available data and models on these issues are not yet robust enough to apply to the sort of exercise here depicted, particularly in the case of Sahara-Sahel (Brito et al., 2014). The accuracy of climate models depends on the distribution of weather stations across the landscape, which is usually low in regions with low human population density and economically unprivileged countries/regions (Bachelet et al., 2016). We attempt to compensate this issue by using a large pixel size, but even considering this caveat our maps constitute first steps to better understand the climate change impacts on tourism in African deserts and the African continent as a whole (Hoogendoorn & Fitchett, 2018). Lastly, we portrayed uncertainty in climate projections in our maps, an improvement over previous studies. Looking into how tourism-dependent communities will be affected by climate-induced changes is an important area of future research (Kaján & Saarinen, 2013). Understanding the perceptions and reactions of all key tourism stakeholders to climate change will be critical for anticipating geographic and seasonal shifts in tourism demand, changes to tourism marketing and overall competitiveness of destinations (Gössling et al., 2006, 2012; Hambira & Saarinen, 2015; Saarinen et al., 2012; Saarinen & Tervo, 2006; Tervo-Kankare et al., 2017). There is an overall need for an increased research into the impacts of climate change on African tourism (Hoogendoorn & Fitchett, 2018) and we support this urgent call. Several research themes on tourism climatology have been suggested by Freitas (2017), some of which are timely relevant for Sahara-Sahel and other drylands facing increased tourism risks from climate change.

Climate impacts on Sahara-Sahel tourism hotspots Our results show that most of Sahara-Sahel will become too hot for tourism, especially during the peak summer seasons. These data-based findings support previous discussions that large parts of northern and southern Africa will become generally less suitable for tourists (see Dube & Nhamo, 2018; Hoogendoorn & Fitchett, 2018). Our results for the warmest months are also consistent with the findings of other studies where substantial impacts on tourism are predicted for the hot seasonal peaks (e.g. Hein et al., 2009 and Köberl et al., 2016). Precipitation projections for deserts vary widely among models (Bachelet et al., 2016), which may somehow explain the irregular results observed for future differences in precipitation along the study area. However, opposite rainfall trends on different areas were found in other African regions where tourism is a substantial contributor to the regional economies (e.g. Dube & Nhamo, 2018). Drastically different temperature and precipitation regimes will probably impact desert-adapted fauna and flora (Brito & Pleguezuelos, 2020; Vale & Brito, 2015), ecosystems

240 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

(Dinerstein et al., 2017), and cultural heritage (Brooks et al., 2020; Winkler & Brooks, 2020), with subsequent implications for client demand and tourism stakeholders (Fisichelli et al., 2015; Gössling & Hall, 2006). Dissimilarities between the two SSPs were expected, as the ‘business-as-usual’ SSP585 points out more pronounced climate change effects than more sustainable paths like the SSP126 (Riahi et al., 2017). Indeed, many more tourism hotspots will be threatened if economic growth keeps at the current pace without sustainable options considered. Under the best-case scenario (SSP126), although most of the southwestern tourism hotspots will be affected by increases in mean annual temperature (BIO 01), it seems that extreme temperatures during the warmest months (BIO 05) will not cause large impacts. The strongest impacts will be mostly felt on regions with few or any tourism hotspots between Algeria, Mauritania, and Mali (Santarém et al., 2020b). This is a positive sign since, if the most sustainable path is taken (SSP126), visitors can travel between hotspots without crossing regions more highly affected by heat peaks. On the other hand, in a desert subject to increased climate extremes, even temperatures below 45ºC during the hottest months pose serious risks to human lives and threatens any kind of tourism. Regional patterns in the distribution of tourism hotspots affected by both year-long and seasonally high temperatures indicate that more countries will be more affected under the worst- case scenario (Algeria, Burkina Faso, Chad, Cameroon, Mali, Mauritania, Niger, Nigeria, and Sudan) than under the most optimistic future scenario (Burkina Faso, Mauritania, and Mali). These regional patterns highlight the need for considering contextual socioeconomic conditions for ameliorating future intolerable temperatures, as the economically most unprivileged from the list are also among the most vulnerable and, thus, less prepared to manage climate change risks. Still, these countries need to develop the most urgent measures to cope with extreme temperatures. Although the Sahara-Sahel tourism hotspots are severely threatened by projected climatic shifts in temperature and precipitation, there is no knowledge about climate impacts on North-African heritage sites (including monuments, places, and rock painting) with respect to the capacity of current management systems to deal with climate change consequences over the coming decades. For instance, the rock-art of Libyan Fezzan and the old desert town of Djenné (Mali) are particularly vulnerable and endangered by changes in heating and precipitation and may lose touristic attractiveness if climate impacts are not rapidly addressed (Brooks et al., 2020). Future policies will be needed to adapt to climate shifts in the region (Brito et al., 2014; Cook & Vizy, 2015; Durant et al., 2014; Taylor et al., 2017; Vale & Brito, 2015).

FCUP 241 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Conclusions: implications for tourism planning, development, and management in Sahara- Sahel Tourism plays a very important economic and social role in the Sahara-Sahel region. In the future, however, extreme droughts in the region will likely bring serious challenges to the economy and human communities who depend heavily on the local climate and weather conditions. Direct impacts on biodiversity, amplified by an increased pressure on natural resources may drive further ecological impacts, deteriorating and disfiguring destinations and keeping away tourists specialized in desert attractions (Saarinen et al., 2012). Even in the best- case scenario, climate projections indicate the region will experience a clear increase in tourism seasonality. Sectorial-specialised tourists (e.g. desert-tourists or ecotourists; see Santarém et al., 2020a and Weaver, 2005) might lose interest in desert attractions if climate become intolerable in Sahara-Sahel under the climate scenarios here studied. This is especially relevant for countries such as Mauritania, where local tourism hotspots have been proposed for differential market segments (e.g. mountain rock-pools for hard-ecotourists and river basins for soft-ecotourists; Santarém et al., 2018), but where warmer climate will threat the long-term sustainable development within those hotspots in the future. Furthermore, this is also the future reality for other sectors such as hospitality and food services, especially when considering the predicted negative impacts on the societal infrastructure need for the tourism industry. Tourism is increasingly considered as a potential driver for poverty alleviation (United Nations, 2018; World Tourism Organization, 2002). Yet, an inclusive and sustainable development for poor, vulnerable and marginalised people will be challenged by climate change, as it is projected to increase the displacement of people and create conflicts related to resource uses (IPCC, 2018). Amplified insecurity will be particularly challenging for many North African regions where conflicts are already impacting local people and biodiversity and are predicted to be exacerbated by global warming (Brito et al., 2018). Poor socio-political contexts in most of the Sahara-Sahel countries negatively influence the likelihood of strong tourism-climate policy integration (Becken et al., 2020). As low-income countries are more vulnerable and less resilient to climate change, they will need to improve social conditions and political stability to become more resilient to climate change (Becken, 2005; Dillimono & Dickinson, 2015; Dogru et al., 2019). In African context, the tourism sector has not advanced much in preparing for the impacts of climate change (Hambira et al., 2020). Tourism stakeholders are often reactive in adaptation and mitigation (Tervo-Kankare et al., 2018) and they are likely overestimating their ability to adapt to higher-emission scenarios and most governments have not undertaken strategies within the

242 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

tourism sector for combating global warming (Scott & Becken, 2010). Among global tourism leaders, there are disagreements on the implementation of climate policies to curb climate change and measurable progresses towards emission reductions (Gössling & Scott, 2018). For countries within Sahara-Sahel, only Egypt, Morocco, and Tunisia have committed strong policies covering the tourism-climate change nexus (Becken et al., 2020). Strong leadership will be needed to reduce climate change risks for tourism in the long run (Gössling & Higham, 2020; Gössling & Scott, 2018) and tourism policymakers and destination management organisations will need to manifestly coordinate integrative climate change policies that co-benefit all tourism stakeholders and sectors in Sahara-Sahel (Becken et al., 2020; UNESCO, 2003). Adaptation measures will be needed to cope with climate change impacts at macro- and micro-scales in the Sahara-Sahel (Becken, 2005; Dogru et al., 2019; Hambira & Saarinen, 2015; Kaján & Saarinen, 2013). Environmental education to local communities and businesses that stress the importance of water conservation will be key for adaptation plans (Saarinen et al., 2012), and traditional architecture adapted for the desert environment should be utilised in future rural and especially urban planning. Tourism businesses in the most impacted hotspots also need to seriously consider their core products in the future (see Hambira et al., 2020). Related to product development, tourist activities based on astronomy and indigenous culture, arts and gastronomy, for example, could be key in desert areas subjected to severe climate change impacts but where alternatives exist to provide unique experiences to tourists (Hambira & Saarinen, 2015; Marshall et al., 2011; Santarém et al., 2020a). Another important strategy for climate change risk management is to map regional tourism infrastructures and asses the risk of various hazards on different locations (Becken, 2005). Still, adaptation measures need to safeguard basic human well-being and rights before addressing the resource needs and challenges of tourism-related businesses (Hoogendoorn & Fitchett, 2018), as otherwise competing and conflicting interests may evolve between local communities and the tourism industry. Therefore, vulnerability assessments and further improvements to adapt to climate change call for stakeholder participation. This research focused on foreseeable impacts of climate change. Although we did not explore adaptation measures per se, the study serves as baseline for the much-needed research in those fronts. When the global costs of climate change are predicted to be in the range of USD486-563 trillion (Calel et al., 2020), the global tourism industry and stakeholders will need to tackle urgent measures to deal with the greatest threat to Humanity. This work addressed the need for understanding the vulnerability of tourism hotspots to climate change on a region that is already challenged by extreme climacteric conditions. Threat maps and projections were obtained FCUP 243 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel from the most refined climate data and with differential socioeconomic pathways for future development. Our results can help regional and international policymakers and tourism planners and managers to plan adaptation strategies for the dramatic predicted impact on tourism hotspots in Sahara-Sahel. The presented framework can be used to evaluate climatic resilience of particular tourism hotspots with strong implications for tourism planning, development, and management. Furthermore, the methodology is easily replicable also to other areas where sustainable tourism is a strong driver of the economy and social development.

Acknowledgements Financial support was given by the European Union’s Horizon 2020 Research and Innovation Programme (Grant 857251), and by the Portuguese Fundação para a Ciência e Tecnologia (FCT) through Strategic Funding UIDB/04423/2020 and UIDP/04423/2020. F.S. and J.C.B. are supported by FCT through grant PD/BD/132407/2017 and contract CEECINST/00014/2018; J.S. is supported by Academy of Finland through grant 272168. We acknowledge the World Climate Research Programme, which, through its Working Group on Coupled Modelling, coordinated and promoted CMIP6. We thank the climate modeling groups for producing and making available their model output, the Earth System Grid Federation (ESGF) for archiving the data and providing access, and the multiple funding agencies who support CMIP6 and ESGF projects. .

244 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

References Amelung, B., & Nicholls, S. (2014). Implications of climate change for tourism in Australia. Tourism Management, 41, 228–244. https://doi.org/10.1016/j.tourman.2013.10.002 Amelung, B., Nicholls, S., & Viner, D. (2007). Implications of global climate change for tourism flows and seasonality. Journal of Travel Research, 45(3), 285–296. https://doi.org/10.1177/0047287506295937 Bachelet, D., Ferschweiler, K., Sheehan, T., & Strittholt, J. (2016). Climate change effects on southern California deserts. Journal of Arid Environments, 127, 17–29. https://doi.org/10.1016/j.jaridenv.2015.10.003 Becken, S. (2005). Harmonising climate change adaptation and mitigation: The case of tourist resorts in Fiji. Global Environmental Change, 15(4), 381–393. https://doi.org/10.1016/j.gloenvcha.2005.08.001 Becken, S. (2013). A review of tourism and climate change as an evolving knowledge domain. Tourism Management Perspectives, 6, 53–62. https://doi.org/10.1016/j.tmp.2012.11.006 Becken, S., Whittlesea, E., Loehr, J., & Scott, D. (2020). Tourism and climate change: evaluating the extent of policy integration. Journal of Sustainable Tourism, 28(10), 1603–1624. https://doi.org/10.1080/09669582.2020.1745217 Becken, S., Zammit, C., & Hendrikx, J. (2015). Developing Climate Change Maps for Tourism: Essential Information or Awareness Raising? Journal of Travel Research, 54(4), 430–441. https://doi.org/10.1177/0047287514528286 Bosello, F., Roson, R., & Tol, R. S. J. (2006). Economy-wide estimates of the implications of climate change: Human health. Ecological Economics, 58(3), 579–591. https://doi.org/10.1016/j.ecolecon.2005.07.032 Boucher, O., Servonnat, J., Albright, A. L., Aumont, O., Balkanski, Y., Bastrikov, V., Bekki, S., Bonnet, R., Bony, S., Bopp, L., Braconnot, P., Brockmann, P., Cadule, P., Caubel, A., Cheruy, F., Codron, F., Cozic, A., Cugnet, D., D’Andrea, F., … Vuichard, N. (2020). Presentation and Evaluation of the IPSL-CM6A-LR Climate Model. Journal of Advances in Modeling Earth Systems, 12(7), 1–52. https://doi.org/10.1029/2019MS002010 Brito, J. C., Durant, S. M., Pettorelli, N., Newby, J., Canney, S., Algadafi, W., Rabeil, T., Crochet, P.-A., Pleguezuelos, J. M., Wacher, T., Smet, K. De, Gonçalves, D. V., da Silva, M. J. F., Martínez-Freiría, F., Abáigar, T., Campos, J. C., Comizzoli, P., Fahd, S., Fellous, A., … Carvalho, S. B. (2018). Armed conflicts and wildlife decline : Challenges and recommendations for effective conservation policy in the Sahara-Sahel. Conservation Letters, 11, 1–13. https://doi.org/10.1111/conl.12446 FCUP 245 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Brito, J. C., Godinho, R., Martínez-Freiría, F., Pleguezuelos, J. M., Rebelo, H., Santos, X., Vale, C. G., Velo-Antón, G., Boratyński, Z., Carvalho, S. B., Ferreira, S., Gonçalves, D. V., Silva, T. L., Tarroso, P., Campos, J. C., Leite, J. V., Nogueira, J., Alvares, F., Sillero, N., … Carranza, S. (2014). Unravelling biodiversity, evolution and threats to conservation in the sahara-sahel. Biological Reviews, 89(1), 215–231. https://doi.org/10.1111/brv.12049 Brito, J. C., & Naia, M. (2020). Coping with Sea-Level Rise in African Protected Areas: Priorities for Action and Adaptation Measures JOSÉ. BioScience, 70(10), 924–932. https://doi.org/10.1093/biosci/biaa087 Brito, J. C., & Pleguezuelos, J. M. (2020). Desert Biodiversity—World’s Hot Spots/Globally Outstanding Biodiverse Deserts. In Encyclopedia of the World’s Biomes. Elsevier Inc. https://doi.org/10.1016/b978-0-12-409548-9.11794-4 Brito, J. C., Tarroso, P., Vale, C. G., Martínez-Freiría, F., Boratyński, Z., Campos, J. C., Ferreira, S., Godinho, R., Gonçalves, D. V., Leite, J. V., Lima, V. O., Pereira, P., Santos, X., da Silva, M. J. F. . F., Silva, T. L., Velo-Antón, G., Veríssimo, J., Crochet, P. A., Pleguezuelos, J. M., & Carvalho, S. B. (2016). Conservation biogeography of the Sahara-Sahel: Additional protected areas are needed to secure unique biodiversity. Diversity and Distributions, 22(4), 371–384. https://doi.org/10.1111/ddi.12416 Brooks, N., Clarke, J., Ngaruiya, G. W., & Wangui, E. E. (2020). African heritage in a changing climate. Azania, 1–32. https://doi.org/10.1080/0067270X.2020.1792177 Bynum, G. D., Pandolf, K. B., & Schuette, W. H. (1978). Induced hyperthermia in sedated humans and the concept of critical thermal maximum. American Journal of Physiology - Regulatory Integrative and Comparative Physiology, 4(3). Cherlet, M., Hutchinson, C., Reynolds, J., Hil, J., Sommer, S., & von Maltitz, G. (2018). World Atlas of Desertification (3th ed.). Publication Office of the European Union. https://doi.org/10.2760/9205 Cook, K. H., & Vizy, E. K. (2015). Detection and analysis of an amplified warming of the Sahara Desert. Journal of Climate, 28(16), 6560–6580. https://doi.org/10.1175/JCLI-D-14-00230.1 Couvreur, T. L. P., Dauby, G., Blach-Overgaard, A., Deblauwe, V., Dessein, S., Droissart, V., Hardy, O. J., Harris, D. J., Janssens, S. B., Ley, A. C., Mackinder, B. A., Sonké, B., Sosef, M. S. M., Stévart, T., Svenning, J. C., Wieringa, J. J., Faye, A., Missoup, A. D., Tolley, K. A., … Sepulchre, P. (2020). Tectonics, climate and the diversification of the tropical African terrestrial flora and fauna. Biological Reviews, 593. https://doi.org/10.1111/brv.12644 Daskin, J. H., & Pringle, R. M. (2018). Warfare and wildlife declines in Africa’s protected areas. Nature, 553(7688), 328–332. https://doi.org/10.1038/nature25194

246 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Di Marco, M., Ferrier, S., Harwood, T. D., Hoskins, A. J., & Watson, J. E. M. (2019). Wilderness areas halve the extinction risk of terrestrial biodiversity. Nature, 573, 582–585. https://doi.org/https://doi.org/10.1038/s41586-019-1567-7 Dillimono, H. D., & Dickinson, J. E. (2015). Travel, tourism, climate change, and behavioral change: travelers’ perspectives from a developing country, Nigeria. Journal of Sustainable Tourism, 23(3), 437–454. https://doi.org/10.1080/09669582.2014.957212 Dinerstein, E., Olson, D., Joshi, A., Vynne, C., Burgess, N. D., Wikramanayake, E., Hahn, N., Palminteri, S., Hedao, P., Noss, R., Hansen, M., Locke, H., Ellis, E. C., Jones, B., Barber, C. V., Hayes, R., Kormos, C., Martin, V., Crist, E., … Kindt, R. (2017). An ecoregion-based approach to protecting half the terrestrial realm. BioScience, 67(6), 534–545. https://doi.org/10.1093/biosci/bix014 Dioko, L. (Don) A. N., & Harrill, R. (2019). Killed while traveling – Trends in tourism-related mortality, injuries, and leading causes of tourist deaths from published English news reports, 2000–2017 (1H). Tourism Management, 70(August 2018), 103–123. https://doi.org/10.1016/j.tourman.2018.08.002 Dogru, T., Marchio, E. A., Bulut, U., & Suess, C. (2019). Climate change: Vulnerability and resilience of tourism and the entire economy. Tourism Management, 72(December 2018), 292–305. https://doi.org/10.1016/j.tourman.2018.12.010 Dong, B., & Sutton, R. (2015). Dominant role of greenhouse-gas forcing in the recovery of Sahel rainfall. Nature Climate Change, 5(8), 757–760. https://doi.org/10.1038/nclimate2664 Dube, K., & Nhamo, G. (2018). Climate variability, change and potential impacts on tourism: Evidence from the Zambian side of the Victoria Falls. Environmental Science and Policy, 84(January), 113–123. https://doi.org/10.1016/j.envsci.2018.03.009 Dubois, G., & Ceron, J. P. (2006). Tourism and climate change: Proposals for a research agenda. Journal of Sustainable Tourism, 14(4), 399–415. https://doi.org/10.2167/jost539.0 Durant, S. M., Pettorelli, N., Bashir, S., Woodroffe, R., Wacher, T., Ornellas, P. de, Ransom, C., Abdelgadir, M., Alqamy, H. E. L., Beddiaf, M., Belbachir, F., Berbash, A., Boitani, L., Breitenmoser, C., Cano, M., Chardonnet, P., Collen, B., Cornforth, W., Cuzin, F., … Baillie, J. (2012). Forgotten biodiversity in desert ecosystems. Science, 336(6087), 1379–1380. https://doi.org/10.1126/science.336.6087.1379 Durant, S. M., Wacher, T., Bashir, S., Woodroffe, R., De Ornellas, P., Ransom, C., Newby, J., Abáigar, T., Abdelgadir, M., El Alqamy, H., Baillie, J., Beddiaf, M., Belbachir, F., Belbachir- Bazi, A., Berbash, A. A., Bemadjim, N. E., Beudels-Jamar, R., Boitani, L., Breitenmoser, C., … Pettorelli, N. (2014). Fiddling in biodiversity hotspots while deserts burn? Collapse of the FCUP 247 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Sahara’s megafauna. Diversity and Distributions, 20(1), 114–122. https://doi.org/10.1111/ddi.12157 ESRI. (2012). ArcGIS Release 10.1. Redlands. Eyring, V., Bony, S., Meehl, G. A., Senior, C. A., Stevens, B., Stouffer, R. J., & Taylor, K. E. (2016). Overview of the Coupled Model Intercomparison Project Phase 6 (CMIP6) experimental design and organization. Geoscientific Model Development, 9(5), 1937–1958. https://doi.org/10.5194/gmd-9-1937-2016 Fang, Y., Yin, J., & Wu, B. (2017). Climate change and tourism: a scientometric analysis using CiteSpace. Journal of Sustainable Tourism, 26(1), 108–126. https://doi.org/10.1080/09669582.2017.1329310 Fick, S. E., & Hijmans, R. J. (2017). WorldClim 2: new 1-km spatial resolution climate surfaces for global land areas. International Journal of Climatology, 37(12), 4302–4315. https://doi.org/10.1002/joc.5086 Fisichelli, N. A., Schuurman, G. W., Monahan, W. B., & Ziesler, P. S. (2015). Protected area tourism in a changing climate: Will visitation at US National Parks warm up or overheat? PLoS ONE, 10(6), 1–13. https://doi.org/10.1371/journal.pone.0128226 Freitas, C. R. de. (2017). Tourism climatology past and present : A review of the role of the ISB Commission on Climate , Tourism and Recreation. International Journal of Biometereology, 61, 107–114. https://doi.org/10.1007/s00484-017-1389-y Garnett, S. T., Burgess, N. D., Fa, J. E., Fernández-Llamazares, Á., Molnár, Z., Robinson, C. J., Watson, J. E. M., Zander, K. K., Austin, B., Brondizio, E. S., Collier, N. F., Duncan, T., Ellis, E., Geyle, H., Jackson, M. V., Jonas, H., Malmer, P., McGowan, B., Sivongxay, A., & Leiper, I. (2018). A spatial overview of the global importance of Indigenous lands for conservation. Nature Sustainability, 1(7), 369–374. https://doi.org/10.1038/s41893-018-0100-6 Gidden, M. J., Riahi, K., Smith, S. J., Fujimori, S., Luderer, G., Kriegler, E., van Vuuren, D. P., van den Berg, M., Feng, L., Klein, D., Calvin, K., Doelman, J. C., Frank, S., Fricko, O., Harmsen, M., Hasegawa, T., Havlik, P., Hilaire, J., Hoesly, R., … Takahashi, K. (2019). Global emissions pathways under different socioeconomic scenarios for use in CMIP6: a dataset of harmonized emissions trajectories through the end of the century. Geoscientific Model Development, 12(4), 1443–1475. https://doi.org/10.5194/gmd-12-1443-2019 Gössling, S., Bredberg, M., Randow, A., Sandström, E., & Svensson, P. (2006). Tourist perceptions of climate change: A study of international tourists in Zanzibar. Current Issues in Tourism, 9(4–5), 419–435. https://doi.org/10.2167/cit265.0 Gössling, S., & Hall, C. M. (2006). Uncertainties in predicting tourist flows under scenarios of

248 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

climate change. Climatic Change, 79(3–4), 163–173. https://doi.org/10.1007/s10584-006- 9081-y Gössling, S., & Higham, J. (2020). The Low-Carbon Imperative: Destination Management under Urgent Climate Change. Journal of Travel Research. https://doi.org/10.1177/0047287520933679 Gössling, S., & Scott, D. (2018). The decarbonisation impasse: global tourism leaders’ views on climate change mitigation. Journal of Sustainable Tourism, 26(12), 2071–2086. https://doi.org/10.1080/09669582.2018.1529770 Gössling, S., Scott, D., & Hall, C. M. (2020). Pandemics, tourism and global change: a rapid assessment of COVID-19. Journal of Sustainable Tourism, 0(0), 1–20. https://doi.org/10.1080/09669582.2020.1758708 Gössling, S., Scott, D., Hall, C. M., Ceron, J. P., & Dubois, G. (2012). Consumer behaviour and demand response of tourists to climate change. Annals of Tourism Research, 39(1), 36–58. https://doi.org/10.1016/j.annals.2011.11.002 Hajima, T., Watanabe, M., Yamamoto, A., Tatebe, H., Noguchi, M. A., Abe, M., Ohgaito, R., Ito, A., Yamazaki, D., Okajima, H., Ito, A., Takata, K., Ogochi, K., Watanabe, S., & Kawamiya, M. (2020). Development of the MIROC-ES2L Earth system model and the evaluation of biogeochemical processes and feedbacks. Geoscientific Model Development, 13(5), 2197– 2244. https://doi.org/10.5194/gmd-13-2197-2020 Hall, C. M. (2019). Constructing sustainable tourism development: The 2030 agenda and the managerial ecology of sustainable tourism. Journal of Sustainable Tourism, 27(7), 1044– 1060. https://doi.org/10.1080/09669582.2018.1560456 Hall, C. M., Scott, D., & Gössling, S. (2020). Pandemics, transformations and tourism: be careful what you wish for. Tourism Geographies, 22(3), 577–598. https://doi.org/10.1080/14616688.2020.1759131 Hambira, W. L., & Saarinen, J. (2015). Policy-makers’ perceptions of the tourism–climate change nexus: Policy needs and constraints in Botswana. Development Southern Africa, 32(3), 350– 362. https://doi.org/10.1080/0376835X.2015.1010716 Hambira, W. L., Saarinen, J., Manwa, H., & Atlhopheng, J. R. (2013). Climate Change Adaptation Practices in Nature-Based Tourism in Maun in the Okavango Delta Area, Botswana: How Prepared are the Tourism Businesses? Tourism Review International, 17(1), 19–29. https://doi.org/10.3727/154427213x13649094288025 Hambira, W. L., Saarinen, J., & Moses, O. (2020). Climate change policy in a world of uncertainty: changing environment, knowledge, and tourism in Botswana. African Geographical Review, FCUP 249 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

39(3), 252–266. https://doi.org/10.1080/19376812.2020.1719366 Hein, L., Metzger, M. J., & Moreno, A. (2009). Potential impacts of climate change on tourism; a case study for Spain. Current Opinion in Environmnetal Sustainability, 1, 170–178. https://doi.org/10.1016/j.cosust.2009.10.011 Hewer, M. J., & Gough, W. A. (2018). Thirty years of assessing the impacts of climate change on outdoor recreation and tourism in Canada. Tourism Management Perspectives, 26, 179– 192. https://doi.org/10.1016/j.tmp.2017.07.003 Hoogendoorn, G., & Fitchett, J. M. (2018). Tourism and climate change: a review of threats and adaptation strategies for Africa. Current Issues in Tourism, 21(7), 742–759. https://doi.org/10.1080/13683500.2016.1188893 Hosni, E. (2000). Strategy for Sustainable Tourism Development in the Sahara. In UNESCO (Issue 32). http://www.unesco.org/new/en/culture/themes/dynamic-content-single- view/news/strategy_for_sustainable_tourism_development_in_the_sahara/ IPCC. (2018). Summary for Policymakers. In V. Masson-Delmotte, P. Zhai, H.-O. Pörtner, D. Roberts, J. Skea, P. R. Shukla, A. Pirani, W. Moufouma-Okia, C. Péan, R. Pidcock, S. Connors, J. B. R. Matthews, Y. Chen, X. Zhou, M. I. Gomis, E. Lonnoy, T. Maycock, M. Tignor, & T. Waterfield (Eds.), Global Warming of 1.5°C (p. 32). World Meteorological Organization. https://report.ipcc.ch/sr15/pdf/sr15_spm_final.pdf%0Ahttp://www.ipcc.ch/report/sr15/ Kaján, E., & Saarinen, J. (2013). Tourism, climate change and adaptation: A review. Current Issues in Tourism, 16(2), 167–195. https://doi.org/10.1080/13683500.2013.774323 Köberl, J., Prettenthaler, F., & Bird, D. N. (2016). Modelling climate change impacts on tourism demand: A comparative study from Sardinia (Italy) and Cap Bon (Tunisia). Science of the Total Environment, 543, 1039–1053. https://doi.org/10.1016/j.scitotenv.2015.03.099 Kriegler, E., Bauer, N., Popp, A., Humpenöder, F., Leimbach, M., Strefler, J., Baumstark, L., Bodirsky, B. L., Hilaire, J., Klein, D., Mouratiadou, I., Weindl, I., Bertram, C., Dietrich, J. P., Luderer, G., Pehl, M., Pietzcker, R., Piontek, F., Lotze-Campen, H., … Edenhofer, O. (2017). Fossil-fueled development (SSP5): An energy and resource intensive scenario for the 21st century. Global Environmental Change, 42, 297–315. https://doi.org/10.1016/j.gloenvcha.2016.05.015 Kriegler, E., Edmonds, J., Hallegatte, S., Ebi, K. L., Kram, T., Riahi, K., Winkler, H., & van Vuuren, D. P. (2014). A new scenario framework for climate change research: The concept of shared climate policy assumptions. Climatic Change, 122(3), 401–414. https://doi.org/10.1007/s10584-013-0971-5

250 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Loarie, S. R., Duffy, P. B., Hamilton, H., Asner, G. P., Field, C. B., & Ackerly, D. D. (2009). The velocity of climate change. Nature, 462(December), 1052–1055. https://doi.org/10.1038/nature08649 Maddison, D. (2001). In search of warmer climates? The impact of climate change on flows of British tourists. Climatic Change, 49(1–2), 193–208. https://doi.org/10.1023/A:1010742511380 Marshall, N. A., Marshall, P. A., Abdulla, A., Rouphael, T., & Ali, A. (2011). Preparing for climate change: Recognising its early impacts through the perceptions of dive tourists and dive operators in the Egyptian red sea. Current Issues in Tourism, 14(6), 507–518. https://doi.org/10.1080/13683500.2010.512075 Monz, C. A., Gutzwiller, K. J., Hausner, V. H., Brunson, M. W., Buckley, R., & Pickering, C. M. (2020). Understanding and managing the interactions of impacts from nature-based recreation and climate change. Ambio. https://doi.org/10.1007/s13280-020-01403-y Mooney, H., Larigauderie, A., Cesario, M., Elmquist, T., Hoegh-Guldberg, O., Lavorel, S., Mace, G. M., Palmer, M., Scholes, R., & Yahara, T. (2009). Biodiversity, climate change, and ecosystem services. Current Opinion in Environmental Sustainability, 1(1), 46–54. https://doi.org/10.1016/j.cosust.2009.07.006 Mora, C., Spirandelli, D., Franklin, E. C., Lynham, J., Kantar, M. B., Miles, W., Smith, C. Z., Freel, K., Moy, J., Louis, L. V., Barba, E. W., Bettinger, K., Frazier, A. G., Colburn IX, J. F., Hanasaki, N., Hawkins, E., Hirabayashi, Y., Knorr, W., Little, C. M., … Hunter, C. L. (2018). Broad threat to humanity from cumulative climate hazards intensified by greenhouse gas emissions. Nature Climate Change, 8(12), 1062–1071. https://doi.org/10.1038/s41558-018- 0315-6 O’Neill, B. C., Tebaldi, C., van Vuuren, D. P., Eyring, V., Friedlingstein, P., Hurtt, G., Knutti, R., Kriegler, E., Lamarque, J.-F., Lowe, J., Meehl, G. A., Moss, R., Riahi, K., & Sanderson, B. M. (2016). The Scenario Model Intercomparison Project (ScenarioMIP) for CMIP6. Geoscientific Model Development, 9(9), 3461–3482. https://doi.org/10.5194/gmd-9-3461- 2016 OECD/SWAC. (2014). An Atlas of the Sahara-Sahel: Geography, Economics and Security. OECD Publishing. https://doi.org/http://dx.doi.org/10.1787/9789264222359-en Ogunrinde, A. T., Oguntunde, P. G., Akinwumiju, A. S., & Fasinmirin, J. T. (2020). Evaluation of the impact of climate change on the characteristics of drought in Sahel Region of Nigeria: 1971–2060. African Geographical Review, 00(00), 1–19. https://doi.org/10.1080/19376812.2020.1814826 FCUP 251 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Powers, R. P., & Jetz, W. (2019). Global habitat loss and extinction risk of terrestrial vertebrates under future land-use-change scenarios. Nature Climate Change, 9(4), 323–329. https://doi.org/10.1038/s41558-019-0406-z Riahi, K., Rao, S., Krey, V., Cho, C., Chirkov, V., Fischer, G., Kindermann, G., Nakicenovic, N., & Rafaj, P. (2011). RCP 8.5-A scenario of comparatively high greenhouse gas emissions. Climatic Change, 109(1), 33–57. https://doi.org/10.1007/s10584-011-0149-y Riahi, K., van Vuuren, D. P., Kriegler, E., Edmonds, J., O’Neill, B. C., Fujimori, S., Bauer, N., Calvin, K., Dellink, R., Fricko, O., Lutz, W., Popp, A., Cuaresma, J. C., KC, S., Leimbach, M., Jiang, L., Kram, T., Rao, S., Emmerling, J., … Tavoni, M. (2017). The Shared Socioeconomic Pathways and their energy, land use, and greenhouse gas emissions implications: An overview. Global Environmental Change, 42, 153–168. https://doi.org/10.1016/j.gloenvcha.2016.05.009 Rogerson, C. M. (2016). Climate change, tourism and local economic development in South Africa. Local Economy, 31(1–2), 322–331. https://doi.org/10.1177/0269094215624354 Rosselló, J., Becken, S., & Santana-Gallego, M. (2020). The effects of natural disasters on international tourism: A global analysis. Tourism Management, 79(April 2019). https://doi.org/10.1016/j.tourman.2020.104080 Saarinen, J. (2021). Sustainable growth in tourism? Rethinking and resetting sustainable tourism for development. In C.M. Hall, L. Lundmark & J. Zhang (Eds), Degrowth and Tourism: New perspectives on tourism entrepreneurship, destinations and policy, pp. 135-151. Routledge, London. Saarinen, J. (2019). What are wilderness areas for? Tourism and political ecologies of wilderness uses and management in the Anthropocene. Journal of Sustainable Tourism, 27(4), 472– 487. https://doi.org/10.1080/09669582.2018.1456543 Saarinen, J., Hambira, W. L., Atlhopheng, J., & Manwa, H. (2012). Tourism industry reaction to climate change in Kgalagadi South District, Botswana. Development Southern Africa, 29(2), 273–285. https://doi.org/10.1080/0376835X.2012.675697 Saarinen, J., & Tervo, K. (2006). Perceptions and adaptation strategies of the tourism industry to climate change: the case of Finnish nature-based tourism entrepreneurs. International Journal of Innovation and Sustainable Development, 1(3), 214. https://doi.org/10.1504/IJISD.2006.012423 Santarém, F., Campos, J. C., Pereira, P., Hamidou, D., Saarinen, J., & Brito, J. C. (2018). Using multivariate statistics to assess ecotourism potential of water-bodies: A case-study in Mauritania. Tourism Management, 67, 34–46.

252 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

https://doi.org/10.1016/j.tourman.2018.01.001 Santarém, F., & Paiva, F. (2015). Conserving desert biodiversity through ecotourism. Tourism Management Perspectives, 16, 176–178. https://doi.org/10.1016/j.tmp.2015.07.016 Santarém, F., Saarinen, J., & Brito, J. C. (2020a). Desert Conservation and Management: Ecotourism. In M. I. Goldstein & D. A. B. T.-E. of the W. B. DellaSala (Eds.), Encyclopedia of the World’s Biomes (pp. 259–273). Elsevier. https://doi.org/https://doi.org/10.1016/B978- 0-12-409548-9.11827-5 Santarém, F., Saarinen, J., & Brito, J. C. (2020b). Mapping and analysing cultural ecosystem services in conflict areas. Ecological Indicators, 110(July 2019), 105943. https://doi.org/10.1016/j.ecolind.2019.105943 Schleussner, C. F., Donges, J. F., Donner, R. V., & Schellnhuber, H. J. (2016). Armed-conflict risks enhanced by climate-related disasters in ethnically fractionalized countries. Proceedings of the National Academy of Sciences of the United States of America, 113(33), 9216–9221. https://doi.org/10.1073/pnas.1601611113 Scott, D. (2011). Why sustainable tourism must address climate change. Journal of Sustainable Tourism, 19(1), 17–34. https://doi.org/10.1080/09669582.2010.539694 Scott, D., & Becken, S. (2010). Adapting to climate change and climate policy: Progress, Problems and potentials. Journal of Sustainable Tourism, 18(3), 283–295. https://doi.org/10.1080/09669581003668540 Scott, D., Gössling, S., & Hall, C. M. (2012). International tourism and climate change. Wiley Interdisciplinary Reviews: Climate Change, 3(3), 213–232. https://doi.org/10.1002/wcc.165 Scott, D., Hall, C. M., & Gössling, S. (2016). A review of the IPCC Fifth Assessment and implications for tourism sector climate resilience and decarbonization. Journal of Sustainable Tourism, 24(1), 8–30. https://doi.org/10.1080/09669582.2015.1062021 Scott, D., Hall, C. M., & Gössling, S. (2019). Global tourism vulnerability to climate change. Annals of Tourism Research, 77(November 2018), 49–61. https://doi.org/10.1016/j.annals.2019.05.007 Scott, D., McBoyle, G., & Schwartzentruber, M. (2004). Climate change and the distribution of climatic resources for tourism in North America. Climate Research, 27(2), 105–117. https://doi.org/10.3354/cr027105 Séférian, R., Nabat, P., Michou, M., Saint-Martin, D., Voldoire, A., Colin, J., Decharme, B., Delire, C., Berthet, S., Chevallier, M., Sénési, S., Franchisteguy, L., Vial, J., Mallet, M., Joetzjer, E., Geoffroy, O., Guérémy, J. F., Moine, M. P., Msadek, R., … Madec, G. (2019). Evaluation of CNRM Earth System Model, CNRM-ESM2-1: Role of Earth System Processes in Present- FCUP 253 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Day and Future Climate. Journal of Advances in Modeling Earth Systems, 11(12), 4182– 4227. https://doi.org/10.1029/2019MS001791 Sherwood, S. C., & Huber, M. (2010). An adaptability limit to climate change due to heat stress. Proceedings of the National Academy of Sciences of the United States of America, 107(21), 9552–9555. https://doi.org/10.1073/pnas.0913352107 Swart, N. C., Cole, J. N. S., Kharin, V. V., Lazare, M., Scinocca, J. F., Gillett, N. P., Anstey, J., Arora, V., Christian, J. R., Hanna, S., Jiao, Y., Lee, W. G., Majaess, F., Saenko, O. A., Seiler, C., Seinen, C., Shao, A., Sigmond, M., Solheim, L., … Winter, B. (2019). The Canadian Earth System Model version 5 (CanESM5.0.3). Geoscientific Model Development, 12(11), 4823– 4873. https://doi.org/10.5194/gmd-12-4823-2019 Tatebe, H., Ogura, T., Nitta, T., Komuro, Y., Ogochi, K., Takemura, T., Sudo, K., Sekiguchi, M., Abe, M., Saito, F., Chikira, M., Watanabe, S., Mori, M., Hirota, N., Kawatani, Y., Mochizuki, T., Yoshimura, K., Takata, K., O&apos;ishi, R., … Kimoto, M. (2018). Description and basic evaluation of simulated mean state, internal variability, and climate sensitivity in MIROC6. Geoscientific Model Development Discussions, 1–92. https://doi.org/10.5194/gmd-2018-155 Taylor, C. M., Belusic, D., Guichard, F., Parker, D. J., Vischel, T., Bock, O., Harris, P. P., Janicot, S., Klein, C., & Panthou, G. (2017). Frequency of extreme Sahelian storms tripled since 1982 in satellite observations. Nature, 544(7651), 475–478. https://doi.org/10.1038/nature22069 Tervo-Kankare, K., Kaján, E., & Saarinen, J. (2018). Costs and benefits of environmental change: tourism industry’s responses in Arctic Finland. Tourism Geographies, 20(2), 202–223. https://doi.org/10.1080/14616688.2017.1375973 Tervo-Kankare, K., Saarinen, J., Kimaro, M. E., & Moswete, N. N. (2017). Nature-based tourism operators’ responses to changing environment and climate in Uis, Namibia. African Geographical Review, 37(3), 273–282. https://doi.org/10.1080/19376812.2017.1286246 UNCTAD. (2017). Tourism for Transformative and Inclusive Growth (Economic D). United Nations. http://unctad.org/en/PublicationsLibrary/aldcafrica2017_en.pdf UNDP. (2019). Human Development Report 2019: Beyond Income, Beyond Averages, Beyond Today - Inequalities in Human Development in the 21st Century (Human Development Report). United Nations Development Programme. https://www.un-ilibrary.org/economic- and-social-development/human-development-report-2019_838f78fd-en UNEP. (2006a). Global Deserts Outlook (E. Ezcurra (ed.)). United Nations Environment Programme. http://www.unep.org/geo/GDOutlook/ UNEP. (2006b). Tourism and Deserts: A Practical Guide to Managing the Social and

254 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Environmental Impacts in the Desert Recreation Sector. United Nations Environment Programme. UNESCO. (2003). The Sahara of Cultures and People: Towards a strategy for the sustainable development of tourism in the Sahara, in the context of combating poverty (Vol. 3, Issue 12). https://doi.org/10.1017/S0003598X00003768 United Nations. (2018). Promotion of sustainable tourism, including ecotourism, for poverty eradication and environment protection : note /. UN,. http://digitallibrary.un.org/record/1639522 UNWTO. (2015). Tourism and the Sustainable Development Goals. In Tourism and the Sustainable Development Goals. https://doi.org/10.18111/9789284417254 UNWTO. (2020). World Tourism Baromoter. In Growth in international tourist arrivals continues to outpace the economy (Vol. 18, Issue 1). Vale, C. G., & Brito, J. C. (2015). Desert-adapted species are vulnerable to climate change: Insights from the warmest region on Earth. Global Ecology and Conservation, 4, 369–379. https://doi.org/10.1016/j.gecco.2015.07.012 van Vuuren, D. P., Stehfest, E., den Elzen, M. G. J., Kram, T., van Vliet, J., Deetman, S., Isaac, M., Goldewijk, K. K., Hof, A., Beltran, A. M., Oostenrijk, R., & van Ruijven, B. (2011). RCP2.6: Exploring the possibility to keep global mean temperature increase below 2°C. Climatic Change, 109(1), 95–116. https://doi.org/10.1007/s10584-011-0152-3 van Vuuren, D. P., Stehfest, E., Gernaat, D. E. H. J., Doelman, J. C., van den Berg, M., Harmsen, M., de Boer, H. S., Bouwman, L. F., Daioglou, V., Edelenbosch, O. Y., Girod, B., Kram, T., Lassaletta, L., Lucas, P. L., van Meijl, H., Müller, C., van Ruijven, B. J., van der Sluis, S., & Tabeau, A. (2017). Energy, land-use and greenhouse gas emissions trajectories under a green growth paradigm. Global Environmental Change, 42, 237–250. https://doi.org/10.1016/j.gloenvcha.2016.05.008 Voldoire, A., Saint-Martin, D., Sénési, S., Decharme, B., Alias, A., Chevallier, M., Colin, J., Guérémy, J. F., Michou, M., Moine, M. P., Nabat, P., Roehrig, R., Salas y Mélia, D., Séférian, R., Valcke, S., Beau, I., Belamari, S., Berthet, S., Cassou, C., … Waldman, R. (2019). Evaluation of CMIP6 DECK Experiments With CNRM-CM6-1. Journal of Advances in Modeling Earth Systems, 11(7), 2177–2213. https://doi.org/10.1029/2019MS001683 Waldron, A., Mooers, A. O., Miller, D. C., Nibbelink, N., Redding, D., & Kuhn, T. S. (2013). Targeting global conservation funding to limit immediate biodiversity declines. Proceedings of the National Academy of Sciences USA, 110(29), 1–5. https://doi.org/10.5061/dryad.p69t1 Walther, O. J. (2017). Wars and Conflicts in the Sahara-Sahel. In West African Papers (10, Issue FCUP 255 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

10). https://www.oecd-ilibrary.org/content/paper/8bbc5813-en Watson, J. E. M., Venter, O., Lee, J., Jones, K. R., Robinson, J. G., Possingham, H. P., & Allan, J. R. (2018). Protect the last of the wild. Nature, 563, 27–30. https://doi.org/10.1038/ d41586- 018-07183-6 Weaver, D. B. (2001). The encyclopedia of ecotourism. CABI. Weaver, D. B. (2005). Comprehensive and minimalist dimensions of ecotourism. Annals of Tourism Research, 32(2), 439–455. https://doi.org/10.1016/j.annals.2004.08.003 Winkler, D. E., & Brooks, E. (2020). Tracing Extremes across Iconic Desert Landscapes: Socio- Ecological and Cultural Responses to Climate Change, Water Scarcity, and Wildflower Superblooms. Human Ecology, 48(2), 211–223. https://doi.org/10.1007/s10745-020-00145- 5 World Bank. (2010). Economics and Adaptation to Climate Change - Synthesis Report. In Climate Change Policies: Global Challenges and Future Prospects. https://doi.org/10.4337/9781781000885.00012 World Tourism Organization. (2002). Tourism and Poverty Alleviation. World Tourism Organization. WTTC. (2020). Travel & tourism. https://wttc.org/Research/Economic-Impact Wu, T., Lu, Y., Fang, Y., Xin, X., Li, L., Li, W., Jie, W., Zhang, J., Liu, Y., Zhang, L., Zhang, F., Zhang, Y., Wu, F., Li, J., Chu, M., Wang, Z., Shi, X., Liu, X., Wei, M., … Liu, X. (2019). The Beijing Climate Center Climate System Model (BCC-CSM): The main progress from CMIP5 to CMIP6. Geoscientific Model Development, 12(4), 1573–1600. https://doi.org/10.5194/gmd-12-1573-2019 Yukimoto, S., Kawai, H., Koshiro, T., Oshima, N., Yoshida, K., Urakawa, S., Tsujino, H., Deushi, M., Tanaka, T., Hosaka, M., Yabu, S., Yoshimura, H., Shindo, E., Mizuta, R., Obata, A., Adachi, Y., & Ishii, M. (2019). The meteorological research institute Earth system model version 2.0, MRI-ESM2.0: Description and basic evaluation of the physical component. Journal of the Meteorological Society of Japan, 97(5), 931–965. https://doi.org/10.2151/jmsj.2019-051

256 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Chapter 6

General Discussion and Future Perspectives FCUP 257 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

6. General Discussion

The major objective of this thesis was to examine if ecotourism constitutes a sustainable solution in remote deserts subjected to political instability and regional conflicts and exhibiting rare and threatened fauna. Four main goals were targeted for answering this central objective: 1) understand which groups of flagship species could be used in marketing and ecotourism campaigns; 2) identify the most suitable areas for ecotourism development at multiple scales according to ecological, cultural and constraint features; 3) find the performance of countries in preserving and promoting cultural services within Sahara-Sahel ecoregions, given the spatial variability of ecological, cultural and constraints features; and 4) understand the impacts of climate change on the identified tourism hotspots and which places will be unreliable for human survival. The first part of this chapter presents and discusses the major achievements of this thesis concerning each of the abovementioned goals. Prospects for future research on desert ecotourism are then provided. Finally, concluding remarks are presented.

6.1. Key findings

6.1.1. Ecotourism in deserts The production of a review on the current ecotourism research in deserts, desert attractions and constraints to develop ecotourism, and the positive and negative impacts of ecotourism on different desert facets (ecology, economy and sociology) (Chapter 1; Article I) was crucial as, among the large extent of ecotourism research produced to date, very few are devoted exclusively to desert ecosystems (Santarém and Paiva, 2015). This lack of research on desert ecotourism adds to the profound society neglect for desert ecosystems (Durant et al., 2012; Waldron et al., 2013). As deserts contain some of the most threatened biodiversity, fragile environments, and vulnerable cultures in the world (UNEP, 2006b; Brito et al., 2014, 2016, 2018, 2019; Vale et al., 2015), and ecotourism is advocated as a potential complementary tool to preserve the natural and cultural heritage of deserts while improving local livelihoods (Santarém and Paiva, 2015), a review of this nature was timely needed. The review allowed to settle a comprehensive knowledge of ecotourism research in deserts and the basis for the detailed analyses developed in the other sections of this thesis (Chapters 4 to 6). Deserts exhibit several elements that may motivate tourists to visit these areas, from the unique landscapes and geological features to biodiversity that can be found nowhere

258 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

else, passing through historical sites evidencing human occupation that are easily visited. But some deserts, particularly in the Sahara-Sahel ecoregions, also exhibit constraints for ecotourism development, such as natural resources overexploitation and geopolitical conflicts that threat biodiversity and people alike. Deserts contain a wide spectrum of potentials and constraints for ecotourism development and management, which, in turn, may deliver different levels of environmental, economic, and sociocultural impacts across desert ecosystems. As such, ecotourism should be developed in ways that are based on effective management tools. Research about desert ecotourism needs to contemplate the most fine and updated tools (such as decision-support and prioritizing tools) to come up with balanced and refined options for sustainable development.

6.1.2. Flagship species and fleets for ecotourism and conservation in Sahara-Sahel An achievement of this thesis was the development of a new methodology to identify and map flagship species and fleets (Chapter 3; Article II). This helped to solve an ongoing issue: the impracticability of performing questionnaires in remote regions to identify charismatic species. The use of questionnaires to identify flagships were the norm to that date (Barua et al., 2012; Verissimo et al., 2009, 2011, 2013, 2017) but a systematic approach that could be applied to remote regions with poor biological knowledge was timely needed. This thesis provided an answer to this issue and a milestone in the field of conservation social science. Due to the paucity of biological data that characterizes the Sahara-Sahel (Brito et al., 2014), this two-step statistical approach allowed to identify flagship fleets in a systematic way and avoided biases inherent to people’s responses to questionnaires (Home et al., 2009; Veríssimo et al., 2011). To my best knowledge, this was the first study that mapped the distribution of flagship species, which also revealed how poorly key biodiversity in Sahara-Sahel is protected (Brito et al., 2016). Transboundary protected areas are needed to protect imperilled desert animals, some of which may be key to attract donors and tourists able to contribute financially to their conservation and may help improving local poor communities (Fig. 5.1). The identification of flagship species in Sahara-Sahel helped to feed the subsequent works of the Thesis, which needed spatial data related to charismatic desert biodiversity (Chapters 4 and 5). Additionally, it contributes to other studies that need to identify flagships in remote and understudied places or in areas that need flagship data to identify optimal sites for conservation and ecotourism. For example, many international agencies remain on flagship marketing as fundraising tools. Organizations such as the Sahara Conservation Fund (SCF) can FCUP 259 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel find here a support for its works in the field and for publicizing charismatic desert animals in their campaigns.

Fig. 5.1. Examples of flagship species that, due to their morphological and behavioural traits, can attract donors and tourists able to improve local socio-economic conditions in Sahara-Sahel. Photo Credits (clockwise): West-African Crocodile: Frederico Santarém; Addax: Francisco Álvares; Moroccan Spiny-tailed Lizard: Zbyszek Boratyński; and Dorcas Gazelle: Zbyszek Boratyński,

6.1.3. Mapping ecotourism and other cultural ecosystem services 6.1.3.1. Ecotourism potential of water-bodies in Mauritania Evaluating the ecotourism potential of sites is a key issue in tourism management and multiple methodologies have been developed to date to do that. Despite the increased use of GIS tools and participatory methods (such as Delphi methods and Analytic Hierarchy Processes), methodologies independent of subjective criteria and weights were lacking in the literature, which compromises interpretations on the best localities to allocate efforts for ecotourism development. A major achievement of this Thesis was the development of a two-stage statistical approach that combined multi-criteria with ordination and clustering algorithms to access the ecotourism potential of sites in a systematic way (Chapter 4; Article III). To my best knowledge, this was the first time that an ecotourism potential assessment was conducted in a systematic, non-heuristic way that is independent of a-priori variables weighting. The methodology, which was tested in 30 water-bodies of Mauritania that are local hotspots of biodiversity (Vale et al., 2015) and hold rich

260 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

human heritage (Hosni, 2000; UNESCO, 2003), allowed to identify suitable water-bodies for ecotourism development while independently assessing which features are related to ecotourism potential in deserts. The method was also able to group sites for different types of ecotourists (hard and soft; see Weaver 2001, 2005 for definitions), which has implications for ecotourism planners that want to maximize return-on-investment, as they can tailor tourism activities to match the demands of distinct hard- and soft-types of ecotourists. The methodology developed here is scalable to other regions in the world where water-bodies offer opportunities for sustainable development through proper ecotourism planning. Given the local economic and societal development and biodiversity conservation needs, this work provided a ground-breaking basis for sustainable development in Mauritania. The country is historically underfunded for conservation (Waldron et al., 2013), is among the poorest countries (UNDP, 2019) and the least visited in the world (UNWTO, 2019) and lacks infrastructural (WEF, 2017) and health conditions (FAO, 2013). However, the mountains in the south display substantial interest for ecotourism, given the many unique biological features that can be observed there, including flagship species such as the West-African Crocodile (Crocodylus suchus; Fig. 5.2; Tellería, 2008) and the rich historical and cultural facets (Campbell et al., 2006; Hosni, 2000; UNESCO, 2003). If these desert facets are properly explored, substantial social and economic improvements can be expected to the people living near these water-bodies. Ecotourism can help achieve some of the most prominent Sustainable Development Goals for Mauritania, such as the end of poverty (Goal 1), decent work for all (Goal 8) and conservation of terrestrial biodiversity (Goal 15). Still, local and international stakeholders need to line-up strategies and plans for the long-term.

FCUP 261 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. 5.2. The West-African crocodile (Crocodylus suchus), a flagship species inhabiting many water-bodies of southern Mauritania that can be marketed in ecotourism campaigns. The picture was taken in 2016 in one of the lagoons with the largest crocodile population in Mauritania. Nearby, there is a village where local people can benefit socially and economically from ecotourism activities. Photo credits: Frederico Santarém.

6.1.3.2. Cultural ecosystem services in Sahara-Sahel Another achievement of this Thesis was the first assessment of the supply of cultural ecosystem services in the Earth’s largest warm region: the Sahara-Sahel (Chapter 4, Article IV). Ecosystem services mapping has been targeting mostly developed countries (Milcu et al., 2013) and cultural services, in particularly, are the least studied services (Martínez-Harms and Balvanera, 2012; Crossman et al., 2013). By mapping cultural services in Sahara-Sahel, I provided one of the first ecosystem services studies developed in deserts to date. In fact, this was the first study that identified the regions supplying cultural services in Sahara-Sahel. Additionally, this is, to my best knowledge, the first study mapping constraints to cultural services supply. Although the cost-importance of human constraints to ecosystem services (O’Farrell et al., 2010), few studies had included them in mapping exercises to date, nor in deserts nor elsewhere (Martinez-Harms et al., 2015; Hanaček and Rodríguez-Labajos, 2018). In addition to Paracchini et al. (2014), van Zanten et al. (2016) and Komossa et al. (2018), this is one of the first studies covering very large scales (>1,000,000 km2) in cultural services supply mapping. This supports decision-making, by helping to identify priorities for ecosystem management and restoration across political borders and avoids overlooking ecological and social processes that operate across large scales of management (Martínez-Harms et al., 2015; Rosa et al., 2017). This is particularly relevant for the context of Sahara-Sahel, where many countries display different social and economic dynamics and where geo-political conflicts had cross borders for years. A coordinated plan between these countries is needed to protect ecosystems and the benefits they provide to people. Transboundary mega-conservation areas can help sustain key ecosystem services (Brito et al., 2016), but ending long-term conflict near some country-borders is of paramount urgency (Brito et al., 2018) if a unified sustainable development is to be achieved for North-African countries. This work also provided the largest and most complete database on the natural and historical-cultural heritage in Sahara-Sahel to date. Not only did we revealed the enormous number of features in Sahara-Sahel that remain to be explored, but also added to voices concerned to the constant society neglection toward deserts (Durant et al., 2012). While desert ecosystems provide tremendous benefits to people (Davies et al., 2012), as we have shown here, the global society keeps fiddling in biodiversity hotspots while desert biodiversity collapses

262 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

(Durant et al., 2014). I hope that this work overcomes the generalised lack of research interest in studying desert ecosystem services (Lu et al., 2018) and reveals the importance of deserts to people, as their ecosystems supply many services that benefit the Humanity. Recreation is currently very asymmetrical in Sahara-Sahel and not all the attractions are equally harnessed. Cultural tourism is well developed in Morocco and Egypt, for instance, whereas Algeria and Libya display some sites with the largest number of rock-art figures in the world that are unknown to the public. Diverting from cultural tourism to ecotourism is also needed as the latter can contribute to the preservation of deserts, as environmental awareness improves among locals and tourists alike (Santarém et al., 2019). National governments need to open their circle of opportunities and embrace the “Green Economy”, to benefit ecosystems and people alike.

6.1.4. Management of cultural ecosystem services management 6.1.4.1. Conservation planning for cultural services in Sahara-Sahel Another achievement of this Thesis was the evaluation of country performance in preserving and promoting their cultural services (Chapter 5; Article V). This multidisciplinary approach allowed to understand where are located the areas supplying the most cultural services and where threats may limit their supply, as well as which countries within Sahara-Sahel are performing better in terms of ecosystem services management. This was the first study ever conducted for cultural services conservation planning in the region, performed with the most updated and fine data and tools available. Combining meaningful data from biological and social sciences allowed us to provide a first step towards cultural services planning in the Sahara-Sahel. It was showed that there is a high potential among desert and arid lands to supply cultural services. The factors that are most important in attracting ecotourists, which are common across Africa (Hausmann et al., 2017) were also showed. Lastly, it was indicated the differentiated performance of countries in supplying and managing cultural services, which are highly dependent on good governance and peace levels (Bradshaw & Di Minin, 2019). This work provided local governments strategies to improve their socio-economic conditions and visitation restrictions. It identified countries that, despite their supply of cultural services, are missing opportunities for local development. Some countries are receiving 300 times more tourists than others (UNWTO, 2019), despite the cultural services supply discrepancies. Some of the rarest and most charismatic desert animal (like the Nanger dama or the Addax nasomaculatus; Fig. 5.1) persist in some of the least visited countries. These countries urgently FCUP 263 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel need to promote their biodiversity to attract international donors and tourists. That these results can be a call for sustainable local development and conflict amelioration. It was verified that conflicts are highly dynamic in the region, which limit interpretations in the short term. For instance, at the time of the study Burkina Faso stood out as one of the best places in Sahara-Sahel to enjoy cultural services. Yet, in early 2020 the local situation deteriorated, and conflicts arose in the north (France24, 2020). The ecosystems that might attract international tourists are now (at the time of this writing) under severe threat. International cooperation and will from bordering countries are needed to conserve local ecosystems that generate benefits to people and to allow tourists to visit desert landscapes without being under life-threat. Some initiatives are urgently needed to be considered: 1) valuations of the natural capital to highlight the economic importance of desert ecosystems to people (Costanza et al., 2014; Strand et al., 2018); 2) expand the land under protection (Adams et al., 2019; Brito et al., 2016; Pouzols et al., 2014) to ensure the multifunctioning of ecosystems at multiple places and times (Cardinale et al., 2012); 3) strengthen environmental laws to control arms trafficking through international accepted Arms Trade Treaties (Brito et al., 2018); 4) improve governance, transparency and accountability (Díaz et al., 2019; Maron et al., 2019) to enable local societies to be more open to businesses (Brito et al., 2018); 5) promote community-based management of natural resources (Brito et al., 2018; Garnett et al., 2018) to ensure that their expertise is considered in conservation plans (Tilman et al., 2017; Willemen et al., 2020); and 6) develop community-based and flagship-based ecotourism to improve local social and economic conditions and to preserve threatened biodiversity in the long run (Brito et al., 2018; Santarém et al., 2019; UNEP, 2006).

6.1.4.2. Climate change impacts in Sahara-Sahel Evaluating the impacts of climate change was important to understand how tourism managers can deal with the threat (Article VI). When climate change effects are predicted to be faster and stronger in desert ecosystems (Loarie et al. 2009), this Thesis provided the first indication of where impacts will be greater and which tourism hotspots are under severe threat of becoming unreliable for visitation in the future. Despite the extensive literature on tourism-climate change nexus, this was the first exercise ever developed to understand geographically where impacts will need to be thoroughly considered if tourism hotspots are to be visited in the future. By using the most refined and updated climate data and global models under two alternative socioeconomic pathways, this Thesis provided the first assessment of climate change impacts on tourism hotspots. The maps provided are useful as planning tools for tourism and

264 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

highlight future challenges for tourism managers adapt to climate change (Becken et al., 2015). The accuracy of climate models depends on the distribution of weather stations across the landscape (Bachelet et al., 2016); still, these maps are informative at the regional level. They constitute first steps to understand how climate change will impact tourism activities in the largest warm region in the world. There are other bioclimatic data and global climate models that can be used in future studies trying to evaluate the impacts of climate change on tourism and we call researchers to use them comprehensively to improve narratives on climate impacts on tourism worldwide. The work enabled the understanding of the regions where climate shifts will be more pronounced and where human survivability may be challenging. As some areas will be severely impacted through most of the year, it is important to look at how tourism communities are affected by these climate-induced changes and which are the perceptions and reactions of tourism stakeholders to deal with climate change (Kaján & Saarinen, 2013). Additionally, the work provided a call for research into the impacts of climate change on African tourism (Hoogendoorn & Fitchett, 2018). Shifts in temperature and precipitation will defy tourism businesses and managers who need to cope with severe impacts. Adaptation will need to be considered. For example, the timing of tourism activities can be adapted to cooler times of the day to ensure human comfort is maintained (Fisichelli et al., 2015; Hambira et al., 2013) and the nature of activities can be focused on desert astronomy and indigenous culture, foods, and music (Hambira et al., 2020; Hambira & Saarinen, 2015; Marshall et al., 2011; Santarém et al., 2020a).

6.2. Future Prospects

This thesis demonstrated that despite the overall neglect for research and conservation in deserts, these ecosystems provide many cultural services that benefit people and that deserve further consideration for sustainable development. Particularly, the Earth’s largest warm region provides a great case-study for ecotourism research, given its unique natural and cultural features and historical conflicts that threaten local biodiversity and people. The thesis provided a couple of new methods to evaluate ecotourism potential in remote arid regions subject to human constraints at multiple scales. Many of these methods were applied to the desert context for the first time, highlighting the strength of the results here provided. However, there are still challenges to overcome. FCUP 265 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

First, and despite the substantial contribution to the desert ecotourism knowledge that the review (Article I) offered, there are still knowledge gaps to overcome. It would be interesting to perform detailed reviews on the attractions and constraints, as well as the impacts of ecotourism on different desert facets, to each of the World’s deserts. A detailed review of these issues to the Sahara Desert and its contiguous arid region, the Sahel, would be particularly welcomed, given the paucity of knowledge that characterizes the region and the need to improve information for sustainable development (Brito et al., 2014). While this thesis increased the knowledge on the distribution of biological, ecological, cultural and historical features to unprecedent levels (Article IV), with many features compiled and mapped for the first time (e.g., landscape features; see Chapter 5), there are still knowledge gaps in the most under-sampled regions of Sahara-Sahel (e.g. the easternmost part; Siddig, 2014) that will help filling those gaps. However, the huge dimensions of the region (≈11,200,000 km2), and the general remoteness that characterizes the region makes the collection of data a complex task. In addition to that, the increase of insecurity and armed conflicts in the region (Brito et al., 2018; Harmon, 2016; Lanouar and Goaied, 2019; OECD-SWAC, 2014; Weiss, 2016) hinder further field surveys. This issue needs to be tackled if further research is to be developed in the field. There are possible solutions to surpass the lack of ground data, but associated limitations need to be accounted. One possible way is to use free software (e.g., Google Earth) that allow the collection of landscape data useful for assessments of ecotourism potential of sites at different scales. Also, publicly available datasets (such as DesertInfo, 2006; GeoCover, 2018; NGA, 2016; NIMA, 1997) constitute some of the most useful and reliable solutions to overcome the lack of field data for now. In addition, crowd-sourced is a new trend in environmental studies to collect field data without going to the field. For example, by mining data from pictures posted in social media, researchers can now collect substantial amounts of data from their laboratories, without the need to travel long distances with the associated large costs (Hausmann et al., 2017). However, crowd-sourced data has some limitations. First, it is generally collected in the most accessible and visited places, whereas the most remote regions, due to their inaccessibility, remain to be visited by most people and, therefore, few or no data are collected there. Second, data validation remains a key issue, as species (or other features) can be mistaken by other similar ones. Good examples of validation of crowd-sourced data have been arriving from the field of Biology. For instance, nest records of the invasive Vespa velutina are collected in Portugal by citizens and is validated by certified members of the municipalities where the nests are recorded; nests are confirmed first by photographs and then by visiting the place where the nest was recorded (Carvalho et al., 2020). Yet, validating crowd-sourced data in some Sahara-Sahel countries will be challenging, due to

266 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

the general lack of infrastructures and high education that characterizes them (UNDP, 2019). Regional training and capacity building would probably contribute to this overcome this issue. Although in this thesis it was developed a new method to identify flagship species and fleets in a systematic way (Article II), answering researchers’ call for new flagship evaluation methods that avoid biases related to questionnaires (including variables associated with the perceptions of local people about proposed flagships; Home et al., 2009; Veríssimo et al., 2011) may improve flagship identification. For instance, this work identified big mammals (e.g. the African elephant; Loxodonta africana) and venomous snakes (e.g. the Nubian spitting cobra; Naja nubiae) as flagships. Yet, crop-damaging species or species that can kill humans may be difficult to promote as flagships due to their negative symbolic value among local people (Barua et al., 2010). But if the ecological and economic roles of these species can be recognized (for instance, elephant tourism brings USD25 million/annually to African countries; Naidoo et al., 2016), they might be accepted among the people that coexists with them in a daily basis. Also, the recognition of the usefulness of some desert species for medical research (e.g. the use of snake poisonous to heal some health diseases) can raise the profile of some neglected species, thus contributing to attract more funds to their conservation. Considering the motives of both local people and tourists to promote a given species to flagship will support flagship conservation projects further (Takahashi et al., 2012). Supervised machine-learning methods (Kotsiantis, 2007), where models are trained to classify any species as flagship or non-flagship, based on a training set of other world’ flagships, can be an alternative to the methods developed in this thesis. Now that the field of conservation marketing is improving at an accelerated rate, model training to identify future flagships might be possible. Still, knowledge about the species under analyses must be available for effective application (Smith et al., 2012). The maps of flagship hotspots can be improved if more robust distribution data become available or if ecological niche-based models be derived for targeted species. These models have been used for predicting species distribution in the Sahara-Sahel transition zone (Vale et al., 2013) and may be applied to the improve knowledge about flagship species distribution, in Sahara-Sahel or in other remote regions. The effectiveness of using flagships when fundraising for projects to preserve wild places also remains to be studied (McGowan et al., 2020), particularly in desert environments. Lastly, future studies could perform economic valuations of flagship species. Estimates of the tourists’ willingness-to-pay (WTP) to observe and conserve desert flagships will help to understand the subsidies that biodiversity conservation in the region can retrieve and to influence policy makers (Baral et al., 2008; Lindsey et al., 2005). In-site questionnaires to evaluate people’ WTP may be very challenging in a region characterized by remoteness and by the lack of trained staff, but online questionnaires offer a FCUP 267 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel good alternative for places with few tourists. Such studies can give an approximate estimation of the economic value of flagship-based ecotourism, raise awareness among local people towards the benefits of preserving habitats containing these species, and help achieve local sustainable development. This thesis proposed a new method to evaluate the suitability of water-bodies for ecotourism development, which helps in ecotourism site planning in deserts or in other ecosystems (Article III). Despite this work allowed to independently assess which features are related with ecotourism potential and to group sites for different ecotourist demands for the first time, there are some methodological improvements that can be considered to refine its application. First, this work considered only continuous variables to assess water-bodies ecotourism potential, whereas many times assessments of ecotourism potential compile categorical variables. Statistics that explore data dissimilarities and that can jointly analyse continuous and categorical variables, such as Principal Coordinates Analysis (PCoA), can help with this process. Such statistics have been used in ecological studies (e.g. Vale and Brito, 2015), but its use in tourism studies remains unexplored. Second, including economic data into ecotourism assessments will improve our ability to evaluate which water-bodies (and the respective local communities living nearby) need urgent investments for development. Third, this work used crocodiles and birds (as a group) as flagship species for ecotourism assessment at local scale (inland water-bodies), but many other desert flagship exist and may be promoted for conservation and ecotourism in Mauritania. Knowing in which water-bodies these flagship species are located will improve future ecotourism potential assessments. Yet, many aquatic habitats remain under-sampled in Mauritania. Sampling of these habitats have been revealing new species for the country, and updating distribution data for many other species (Santarém et al., 2018). In particular, the West-African crocodile (Crocodylus suchus) has been discovered at many new localities (Campos et al., 2016), offering clues about the under-sampling that characterizes the country. Future surveys should focus on the collection of additional distribution data, especially in the most under-sampled regions, where it is probable to discover additional flagship species populations that will help promote local conservation and ecotourism. In addition, studies need to identify threats affecting aquatic habitats and species (Campos et al., 2016), as this will allow to develop strong conservation plans supported by robust data and that allow the sustainable use of water-bodies by all stakeholders (Vale et al., 2015). Future frameworks should also envisage capacity building and environmental training of local people for conservation and tourism activities (Brito et al., 2018; Sow et al., 2017) (Fig. 5.3). Engaging local communities in conservation and ecotourism planning will contribute to establish solid conservation strategies with long-term success (Brito et

268 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

al., 2018; Garnett et al., 2018; Tilman et al., 2017; Willemen et al., 2020). Above all, any planning needs to be developed in close relation with local people, to understand their willing to engage in tourism activities and to include this alternative land-use in their livelihoods (Weaver, 2005).

Fig. 5.3. Local training of Mauritanian people for conservation and ecotourism. Clockwise: JC Brito, 07/11/2010, Mauritania, Sélibabi; JC Campos, 31/10/2011, Mauritania, Tarf Tazazmout; JC Campos, 08/11/2014, Mauritania, Diawling with Zeine of PND; J Marques, 21/08/2015, Mauritania, with team of PN Diawling

This thesis provided the most complete distribution of biological, conservation, landscape, cultural and historical features to date in Sahara-Sahel (Article IV). In addition, the thesis allowed to understand which regions supply the most cultural services and identify where transboundary constraints threaten desert ecosystems. However, some regions that were not identified as hotpots of cultural services in this study might be supplying them, but the paucity of data that characterizes Sahara-Sahel (Brito et al., 2014) might be dictating a different scenario. The biggest mountains of the study region (e.g. the Tibesti Mountain in Chad) lack high-resolution maps of natural features; yet they might be rich in geological features that add to the ecosystem services value. Despite the remoteness that characterizes the region, social media opens new avenues FCUP 269 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel for data mining to locate cultural ecosystem services in space and time (Oteros-Rozas et al., 2018; Vaz et al., 2019). Still, as discussed before, this data will be only possible to collect by tourists and local people when local conflicts are ameliorated. Range maps at the sub-continental scale, as this thesis provided, provide good estimates to inform biodiversity priorities at large scales, but should ideally be combined with local data and maps before finer-scale conservation decisions are made by governments and conservation agencies (McGowan et al., 2020). Future studies can also consider the demand of desert cultural services (Wolff et al., 2015). Additionally, the monetary value of desert cultural services could be explored in the future. Despite monetary studies of ecosystem services were developed at the global scale (Costanza et al., 2014, 1997), assessments exclusively dedicated to deserts remain to be done (Davies et al., 2012). With this economic data, and in addition to the spatial mapping of cultural services hotspots, one can start thinking of transboundary conservation plans envisioned to protect local ecosystems. To do so, transboundary mega-conservation areas (Brito et al., 2016) and multi-lateral agreements between Sahara-Sahel countries to eliminate conflicts and terrorism (Brito et al., 2018) will be needed. It is urgent that international and local players find solutions that benefit all stakeholders: local communities, governments, NGOs, local companies, and visitors. An initiative on mapping and assessment of ecosystems and their services for the region is timely needed. Such initiatives were done in the European Union (Burkhard and Maes, 2017) but I am unaware of such developments for African nations. Such analytical framework would allow to integrate the value of ecosystems into accounting and reporting systems. This thesis also evaluated countries performance in managing cultural services in Sahara- Sahel (Article V). Still, there are many biological and social issues to be considered for improvement. Biological data are scarce (Brito et al., 2014), as discussed along the Thesis. Although social data are provided to a larger extent than biological data, there are many fields in which improvements need to be made. For instance, tourism data are very inconsistency along Sahara-Sahel countries. Some countries report annual data, whereas others stopped to provide them long ago. Additionally, due to the recent separation between Sudan and South Sudan, tourism data is lacking. The extreme poverty and lack of social infrastructures that characterizes most of these countries (UNDP, 2019; WEF, 2017) detract the robust collection of data that could be useful for research and management across scales. Improving governance and socioeconomic conditions may help to organize better databases for future research, as it has been found elsewhere in Africa (Bradshaw and Di Minin, 2019). Also, some countries urgently need to better explore the opportunities that local ecosystems provide. While some countries keep growing economically due to high tourism investment levels, others are lagging. Research

270 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

specifically developed to highlight the monetary value of local ecosystems is urgent (Costanza et al., 2014), as their contribution to humanity might be much larger than previously thought (Taylor et al., 2017). Improving socioeconomic conditions will be detrimental to attract international tourists able to provide the economic means for conservation and sustainable development in the most remote and neglected regions. Yet, the temporal dynamics of some of the indicators we have used can be accounted in future studies, as they play a key role in prioritizing sites for ecosystem conservation that can quickly change due to anthropogenic factors (Rappaport et al., 2015). This is especially relevant in drylands subject to historical geopolitical conflicts, as it is the case of Sahara-Sahel (Brito et al., 2018). Other constraints, such as sites of natural resources extraction, can be considered in future studies. Overexploitation of resources is a norm in the region (Brito et al., 2014; Duncan, 2014), but such data is detrimental to highlight the costs related to ecosystem services losses. Currently, nature conservation can seem bleak for Sahara-Sahel, but highlighting the value of natural capital (Costanza et al., 2014), expand and properly manage conservation lands (Brito et al., 2016), improve local socioeconomic and governance conditions (Díaz et al., 2019), and enable local knowledge onto conservation planning (Garnett et al., 2018; Willemen et al., 2020) will allow to develop local conditions while improving nature conservation. This thesis also assessed the impacts of climate change on tourism hotspots in the Sahara-Sahel for the first time (Article VI). Yet, some improvements can be made to better understand how to deal with the number one global threat for conservation. For example, we used only two bioclimatic data, eight global climate models and two alternative socio-economic pathways for development. Yet, there are other data to explore and many more will become available in the future (Eyring et al., 2016; Fick & Hijmans, 2017; Rhiai et al., 2017), which will improve model accuracy and results interpretation. There are low-probability events that can have a large influence on travelling, such as terrorism events, natural disasters, and pandemics and which can be explored in future tourism-climate change studies (Gössling et al., 2020; Gössling & Higham, 2020; Hall et al., 2020; Rosselló et al., 2020). Looking at how tourism communities will be affected by these climate-induced changes is also an important area of future research (Kaján & Saarinen, 2013). Although we quantified the impacts of climate change on tourism hotspots providing maps that help tourism managers, understanding the perceptions and reactions of all tourism stakeholders to climate change will be critical for anticipating geographic and seasonal shifts in tourism demand, changes to tourism marketing and overall competitiveness of destinations (Gössling et al., 2012; Hambira & Saarinen, 2015; Saarinen et al., 2012). Lastly, there is a need for an increased research into the impacts of climate change on tourism in North Africa (Hoogendoorn & Fitchett, 2018). Still, tourism policymakers will need to integrate strong FCUP 271 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel climate change policies that co-benefit all tourism stakeholders and sectors in Sahara-Sahel (UNESCO, 2003). Vulnerability assessments and further improvements to mitigate and adapt to climate change will need stakeholder participation from multiple disciplines. Finally, and although assessment of tourism contributions to mammals (Buckley et al., 2012), birds (Steven et al., 2013) and amphibians (Morrison et al., 2012) were conducted in the past, such assessments were never performed for reptiles. I think this is timely needed, especially because most of the desert biota is composed by reptiles (Brito and Pleguezuelos, 2019), many of them flagship candidates to be used in conservation and ecotourism campaigns (Santarém et al., 2019; Tellería et al., 2008). Once ecotourism starts to be considered as a sustainable land- use for development in Sahara-Sahel, future studies will need to quantify the ecological, societal, and economic impacts of its development and the effectiveness of management tools available at that time (Buckley et al., 2016; Mossaz et al., 2015). Tracking visitors and local companies’ behaviour patterns will also be needed to be accounted (Buckley, 2009; Weaver, 2001). The relationships and the dynamics between ecotourism business and local communities, protected areas, governments, and all stakeholders need to be studied as well, as well as the financial viability of ecotourism activities need to be studied (Weaver, 2001). Long-term conservation in deserts will require appropriate policy instruments that promote the sustainable use of natural resources. I believe that the promotion of desert cultural ecosystem services, particularly those related to ecotourism, will help combating land-degradation while preserving key desert diversity and local livelihoods. Raising environmental awareness and pride among local communities for the value of desert natural and cultural heritage is timely needed to pressure strong policy changes (Brito and Pleguezuelos, 2019). There are multiple institutions dedicated to desert research and promotion in some of the world’ deserts (e.g. the Nevada Desert Research Center in North America, Desert Knowledge Research Institute in Australia, and the Desert Research Foundation of Namibia). Yet, these facilities are lacking in most Asian and African deserts (but see the Desert Research Center in Egypt) in a time when they are mostly needed mostly because of the unsustainable development predicted for Saharan and Sahelian nations (UNEPb, 2006; Brito et al., 2018). International, national, and local agencies are called to work together towards the development of such infrastructures, which will help gaining further knowledge on desert heritage and refining tools to preserve it. Improvement of regional governance, social development and human rights enforcement laws will be also needed to achieve local targets of the United Nations Sustainable Development Goals (https://www.un.org). For instance, the Sahara Conservation Fund (SCF) has been able to concert actions between local administrations, international environmental authorities and local people to reintroduce the

272 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

flagship Scimitar-horned oryx (Oryx dammah) in the Ouadi Rimé-Ouadi Achim Game Reserve in Chad, after decades of being considered extinct in the wild (Durant et al., 2014). Similar programs are being created for the reintroduction of the African Ostrich (Struthio camelus) in Niger (Mamadou et al., 2020), and of the mhorr gazelle (Nanger dama mhorr) in Morocco (Abáigar et al., 2019), all of which are flagship species for conservation and ecotourism (Santarém et al., 2019). These are good news and well-established projects that can (and should) be extended to other Saharan countries to save the local biota (e.g., the West-African crocodile, Crocodylus suchus, in Mauritania; Telleria, 2008). Preserving these flagships in similar projects may allow attracting funds and ecotourists able to support further the conservation of charismatic species and associated habitats, as well as the socioeconomic development of the people sharing flagship’ habitats. FCUP 273 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

6.3 Concluding remarks

The major achievements of this thesis provide relevant insights about the importance of ecotourism to desert heritage and how novel planning and statistical tools can help assessments of flagships and of ecotourism potential at various scales. This thesis outlined strengths and weaknesses of available methods for ecotourism planning and provided hints for systematic evaluations. The thesis also provided novel and the most complete spatial data on desert facets for Sahara-Sahel to date, which highly contributed to the knowledge on a region that has been poorly studied and neglected by the society. The interdisciplinary approach developed along the chapters of this thesis provided a strong framework for the sustainable development in Sahara- Sahel countries, which can benefit local biodiversity, ecosystems, and people. The major conclusions of this thesis are presented below: • Future ecotourism studies should focus on the development of integrative frameworks that combine data from ecological, social, and economic fields to better understand the relationships between these variables at different spatial scales. • Multivariate statistics allow robust assessments of the variables that contribute the most to assess flagship species among several candidates and that influence the ecotourism potential of sites. • Increased efforts on the development of statistical, GIS, and conservation planning tools will allow robust assessments of the places supplying the highest levels of CES and where the development of ecotourism will provide the highest return-on-investment to all stakeholders. • Animals’ morphological and behavioural traits, as well as body traits and conservation status are relevant when identifying flagship species (and fleets) for conservation and ecotourism and should be considered when developing plans that target flagship candidates at local scales. • Hotspots of flagship species are coincident with hotspots of general biodiversity in Sahara- Sahel and most of them currently lack proper protection. • Water availability and temperature extremes play a key role in ecotourism planning at many scales (Mauritanian water-bodies and Sahara-Sahel limits), and need to be accounted when designing ecotourism activities. • Different groups of water-bodies are differently suitable for soft and hard ecotourists and local stakeholders should consider these differences when designing tourism activities.

274 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

• Geopolitical conflicts and overexploitation of natural resources constrain the supply of CES and threaten the ecosystem supply of benefits to people in the long run, urging plans to ease ethnic, religious and cultural tensions and to explore natural resources in a sustainable way by adopting the principles of the Green Economy. • Costs related to constraints to ecosystem conservation should be incorporated in conservation planning, in deserts or elsewhere, as it provides a more robust response to where to spend limited conservation efforts and money efficiently. • Despite poorly protected, the main mountains and water-bodies of Sahara-Sahel concentrate hotspots of natural and cultural heritage and are priorities for biological and CES conservation, urging the development of planning schemes that properly protect these regions from unsustainable human development or activities. • Refined conservation planning tools allow to integrate biological and social data in novel ways, providing the best answers for desert ecosystem management. • Tourism development is very asymmetrical across Sahara-Sahel countries, with some countries accounting for most of the tourism arrivals whereas others are missing tourism opportunities for local development. • Cultural tourism is quite well developed in some north-African countries, but ecotourism is still in its infancy. • Governance and socioeconomic conditions play a key role in defining countries ability to conserve and manage local ecosystems. • Climate change is predicted to severely affect tourism hotspots in Sahara-Sahel. • Climate shifts will pose a severe threat to human survivability in deserts. • Managers will need to adapt to climate shifts in the future if tourism is desired to keep contributing to the socioeconomic development of the region • Substantial research needs to be performed in Sahara-Sahel if local people and biodiversity are to be seriously considered by local governments and if sustainable development is the ultimate objective of the region. • The monetary value of desert ecosystems may be much larger than previously thought, which may open new avenues for the world society recognize the value of drylands. • Conservation strategies need to account local needs to develop plans that benefit all: biodiversity, local agencies and governments, and local people. FCUP 275 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

References Abáigar, T., Rodríguez-Caballero, E., Martínez, C., Amaouch, Z., Samlali, M. L., Aparicio, F., El Balla, T., Essalhi, A., Fernández, J., García, F., Haya, M., M’Bareck, A., M’Bareck, H., González, L. M., Fernández de Larrínoa, P. (2019). The first reintroduction project for mhorr gazelle (Nanger dama mhorr) into the wild: Knowledge and experience gained to support future conservation actions. Global Ecology and Conservation, e00680. doi: 10.1016/j.gecco.2019.e00680. Adams, V. M., Iacona, G. D. & Possingham, H. P. (2019). Weighing the benefits of expanding protected areas versus managing existing ones. Nature Sustainability, 2, 404–411. doi: https://doi.org/10.1038/s41893-019-0275-5. Bachelet, D., Ferschweiler, K., Sheehan, T., & Strittholt, J. (2016). Climate change effects on southern California deserts. Journal of Arid Environments, 127, 17–29. https://doi.org/10.1016/j.jaridenv.2015.10.003. Baral, N., Stern, M., J., Bhattarai, R. (2008). Contingent valuation of ecotourism in Annapurna conservation area, Nepal: Implications for sustainable park finance and local development. Ecological Economics, 66, 2-3. doi: https://doi.org/10.1016/j.ecolecon.2008.02.004. Barua, M., Gurdak., D. J., Ahmed, R. A., & Tamuly, J. (2012). Selecting flagship species for invertebrate conservation. Biodiveristy and Conservation, 21, 1457-1476. doi: 10.1007/s10531-012-0257-7. Barua, M., Tamuly, J., & Ahmed, R. A. (2010). Mutiny or clear sailing? Examining the role of the Asian Elephant as a flagship species. Human Dimensions of Wildlife, 15, 145–160. doi: 10.1080/10871200903536176. Becken, S., Zammit, C., & Hendrikx, J. (2015). Developing Climate Change Maps for Tourism: Essential Information or Awareness Raising? Journal of Travel Research, 54(4), 430–441. https://doi.org/10.1177/0047287514528286. Bradshaw, C. J. A. & Di Minin, E. (2019). Socio-economic predictors of environmental performance among African nations. Scientific Reports, 9, 1–13. doi: https://doi.org/10.1038/s41598-019-45762-3. Brito J.C., Pleguezuelos J.M. (2019). Desert Biodiversity—World’s Hot Spots/Globally Outstanding Biodiverse Deserts. In: Encyclopedia of the World’s Biomes. doi: https://doi.org/10.1016/B978-0-12-409548-9.11794-4. Brito, J. C., Durant, S. N., Pettorelli, N., Newby, J., Canney, S., Algadafi, W., Rabeil, T., Crochet, P.-A., Pleguezuelos, J. M., Wacher, T., de Smet, K., Gonçalves, D. V., Ferreira da Silva, M. J., Martínez-Freiría, F., Abáigar, T., Campos, J. C., Comizzoli, P., Fahd, S., Fellous, A., Malam

276 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Garba, H. H., Hamidou, D., Harouna, A., Hatcha, M. S., Nagy, A., Silva, T. L., Sow, A. S., Vale, C. G., Boratyński, Z., Rebelo, H. & Carvalho, S. B. (2018). Armed conflicts and wildlife decline: challenges and recommendations for effective conservation policy in the Sahara-Sahel. Conservation Letters, 11, e12446. doi: https://doi.org/10.1111/conl.12446. Brito, J. C., Godinho, R., Martínez-Freiría, F., Pleguezuelos, J. M., Rebelo, H., Santos, X., Vale, C. G., Velo-Antón, G., Boratyński, Z., Carvalho, S. B., Ferreira, S., Gonçalves, D. V., Silva, T. L., Tarroso, P., Campos, J. C., Leite, J. V., Nogueira, J., Álvares, F., Sillero, N., Sow, A. S., Fahd, S., Crochet, P.-A. & Carranza, S. (2014). Unravelling biodiversity, evolution and threats to conservation in the Sahara-Sahel. Biological Reviews, 89, 215-231. doi: https://doi.org/10.1111/brv.12049. Brito, J. C., Tarroso, P., Vale, C. G., Martínez-Freiría, F., Boratyński, Z., Campos, J. C., Ferreira, S., Godinho, R., Gonçalves, D. V., Leite, J. V., Lima, V. O., Pereira, P., Santos, X., Ferreira da Silva, M. J., Silva, T. L., Velo-Antón, G., Veríssimo, J., Crochet, P.-A., Pleguezuelos, J. M. & Carvalho, S.B. (2016). Conservation biogeography of the Sahara-Sahel: additional protected areas are needed to secure unique biodiversity. Diversity and Distributions, 22, 371-384. doi: https://doi.org/10.1111/ddi.12416. Buckley, R., Morrison, C., & Castley, J. G. (2016). Net effects of ecotourism on threatened species survival. PLoS ONE, 11, e0147988. doi: https://doi.org/10.1371/journal.pone.0147988. Buckley, R.C., Castley, J.G., Pegas, F.d.V., Mossaz, A.C., Steven, R. (2012) A population accounting approach to assess tourism contributions to conservation of IUCN-Redlisted mammal species. PLoS ONE, 7(9), e44134. doi: 10.1371/journal.pone.0044134. Buckley, R. C. (2009) Ecotourism: Principles and Practices. CAB International, Wallingford, USA. Burkhard B and Maes J (Eds.) (2017) Mapping Ecosystem Services. Pensoft Publishers, Sofia, Bulgaria. Campbell, A., Coulson, D., Challis, S., Keenan, J. (2006). Some Mauritanian rock art sites. Sahara Journal, 17, 143-146. Accessed on 14 August 2020 at www.saharajournal.com. Campos, J. C., Martínez-Freiría, F., Sousa, F. V., Santarém, F., & Brito, J. C. (2016). Update of distribution, habitats, population size, and threat factors for the West African crocodile in Mauritania. Amphibia-Reptilia, 37(3), 325–330. https://doi.org/10.1163/15685381- 00003059. Cardinale, B. J., Duffy, J. E., Gonzales, A., Hooper, D. U., Perrings, C., Venail, P., Nawwani, A., Mace, G. M., Tilman, D., Wardle, D. A., Kinzig, A. P., Daily, G. C., Loreau, M., Grace, J. B., Larigauderie, A., Srivastava, D. S., Naeem, S. (2012). Biodiversity loss and its impact on humanity. Nature, 486, 59–67. doi: https://doi.org/10.1038/nature11148. FCUP 277 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Carvalho, J., Hipólito, D., Santarém, F., Martins, R., Gomes, A., Carmo, P., Rodrigues, R., Grosso‐Silva, J., Fonseca, C. (2020) Patterns of Vespa velutina invasion in Portugal using crowdsourced data. Conservation and Diversity. https://doi.org/10.1111/icad.12418 Costanza, R., de Groot, R., Sutton, P., van der Ploeg, S., Anderson, S.J., Kubiszewski, I., Farber, S., Turner, R.K. (2014). Changes in the global value of ecosystem services. Global Environmental Change, 26, 152–158. doi: 10.1016/j.gloenvcha.2014.04.002. Costanza, R., d’Arge, R., de Groot, R., Farber, S., Grasso, M., Hannon, B., Limburg, K., Naeem, S., Oneill, R.V., Paruelo, J., Raskin, R.G., Sutton, P., van den Belt, M., 1997. The value of the world’s ecosystem services and natural capital. Nature, 387, 253–260. doi: https://doi.org/10.1038/387253a0. Crossman, N.D., Burkhard, B., Nedkov, S., Willemen, L., Petz, K., Palomo, I., Drakou, E.G., Martín-López, B., McPhearson, T., Boyanova, K., Alkemade, R., Egoh, B., Dunbar, M.B., Maes, J. (2013). A blueprint for mapping and modelling ecosystem services. Ecosystem Services, 4, 4–14. doi: 10.1016/j.ecoser.2013.02.001. Davies, J., Poulsen, L., Schulte-Herbrüggen, B., Mackinnon, K., Crawhall, N., Henwood, W.D., Dudley, N., Smith, J., Gudka, M. (2012). Conserving Dryland Biodiveristy. IUCN, Nairobi, Kenya. DesertInfo (2006). Landmines: pictures, maps, GPS + Google Earth Data. DesertInfo Forum. Accessed on 06 September 2018 at http://www.desert-info.ch/desert-info- forum/viewtopic.php?t=1927. Díaz, S., Settele, J., Brondízio, E.S., Ngo, H. T., Agard, J., Arneth, A., Balvanera, P., Brauman, K. A., Butchart, S. H. M., Chan, K. M. A., Garibaldi, L. A., Ichii, K., Liu, J., Subramanian, S. M., Midgley, G. F., Miloslavich, P., Molnár, Z., Obura, D., Pfaff, A., Polasky, S., Pruvis, S., Purvis, A., Razzaque, J., Reyers, B., Chowdhury, R. R., Shin, Y.-J., Visseren-Hamakers, I., Willis, K. J., Zayas, C. N. (2019). Pervasive human-driven decline of life on Earth points to the need for transformative change. Science, 366, eaax3100. doi: 10.1126/science.aax3100. Duncan, C., Kretz, D., Wegmann, M., Rabeil, T., Pettorelli, N. (2014). Oil in the Sahara: mapping anthropogenic threats to Saharan biodiversity from space. Philosophical Transactions of the Royal Society of London B., 369, 20130191. doi: 10.1098/rstb.2013.0191. Durant, S. M., Wacher, T., Bashir, S., Woodroffe, R., De Ornellas, P., Ransom, C., … Pettorelli, N. (2014). Fiddling in biodiversity hotspots while deserts burn? Collapse of the Sahara’s megafauna. Diversity and Distributions, 20, 114–122. doi: 10.1111/ddi.12157. Durant, S. M., Pettorelli, N., Bashir, S., Woodroffe, R., Wacher, T., Ornellas, P. de, … Baillie, J. (2012). Forgotten Biodiversity in Desert Ecosystems. Science, 336(6087), 1379–1380. doi:

278 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

10.1126/science.336.6087.1379. Eyring, V., Bony, S., Meehl, G. A., Senior, C. A., Stevens, B., Stouffer, R. J., & Taylor, K. E. (2016). Overview of the Coupled Model Intercomparison Project Phase 6 (CMIP6) experimental design and organization. Geoscientific Model Development, 9(5), 1937–1958. https://doi.org/10.5194/gmd-9-1937-2016. Fick, S. E., & Hijmans, R. J. (2017). WorldClim 2: new 1-km spatial resolution climate surfaces for global land areas. International Journal of Climatology, 37(12), 4302–4315. https://doi.org/10.1002/joc.5086. Fisichelli, N. A., Schuurman, G. W., Monahan, W. B., & Ziesler, P. S. (2015). Protected area tourism in a changing climate: Will visitation at US National Parks warm up or overheat? PLoS ONE, 10(6), 1–13. https://doi.org/10.1371/journal.pone.0128226. France 24 (2020). ‘Armed terrorists’ stage deadly attack on convoy of traders in Burkina Faso. France 24. Accessed on 15 June 2020 at https://www.france24.com/en/20200530-armed- terrorists-stage-deadly-attack-on-convoy-of-traders-in-burkina-faso. Food and Agricultural Organization of the United Nations (FAO) (2013). Low-Income Food-Deficit Countries (LIFDC) — List for 2013. Accessed on February 2017 at http://www.fao.org/countryprofiles/lifdc/en/FAO 2011. Garnett, S. T., Burgess, N. D., Fa, J. E., Fernández-Llamazares, Á., Molnár, Z., Robinson, C. J., Watson, J. E. M., Zander, K. K., Austin, B., Brondizio, E. S., Colier, N. F., Duncan, T., Ellis, E., Geyle, H., Jackson, M. V., Jonas, H., Malmer, P. McGowan, B., Sivongxay, A., Leiper, I. (2018). A spatial overview of the global importance of Indigenous lands for conservation. Nature Sustainability, 1, 369–374. doi: https://doi.org/10.1038/s41893-018-0100-6. GeoCover (2018). Landsat GeoCover. Global Land Cover Facility. Accessed on 06 September 2018 at http://www.landcover.org/research/portal/geocover/. Gössling, S., Scott, D., & Hall, C. M. (2020). Pandemics, tourism and global change: a rapid assessment of COVID-19. Journal of Sustainable Tourism, 0(0), 1–20. https://doi.org/10.1080/09669582.2020.1758708. Gössling, S., Scott, D., Hall, C. M., Ceron, J. P., & Dubois, G. (2012). Consumer behaviour and demand response of tourists to climate change. Annals of Tourism Research, 39(1), 36–58. https://doi.org/10.1016/j.annals.2011.11.002. Hall, C. M., Scott, D., & Gössling, S. (2020). Pandemics, transformations and tourism: be careful what you wish for. Tourism Geographies, 22(3), 577–598. https://doi.org/10.1080/14616688.2020.1759131. FCUP 279 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Hambira, W. L., & Saarinen, J. (2015). Policy-makers’ perceptions of the tourism–climate change nexus: Policy needs and constraints in Botswana. Development Southern Africa, 32(3), 350– 362. https://doi.org/10.1080/0376835X.2015.1010716. Hambira, W. L., Saarinen, J., Manwa, H., & Atlhopheng, J. R. (2013). Climate Change Adaptation Practices in Nature-Based Tourism in Maun in the Okavango Delta Area, Botswana: How Prepared are the Tourism Businesses? Tourism Review International, 17(1), 19–29. https://doi.org/10.3727/154427213x13649094288025. Hambira, W. L., Saarinen, J., & Moses, O. (2020). Climate change policy in a world of uncertainty: changing environment, knowledge, and tourism in Botswana. African Geographical Review, 39(3), 252–266. https://doi.org/10.1080/19376812.2020.1719366. Hanaček, K., Rodríguez-Labajos, B. (2018). Impacts of land-use and management changes on cultural agroecosystem services and environmental conflicts—A global review. Global Environmental Change, 50, 41–59. doi: 10.1016/j.gloenvcha.2018.02.016. Harmon, S.A. (2016). Terror and Insurgency in the Sahara-Sahel Region: Corruption, Contraband, Jihad and the Mali War of 2012-2013. Routledge, London, UK. Hausmann, A., Toivonen, T., Heikinheimo, V. & Tenkanen, H. (2017). Social media reveal that charismatic species are not the main attractor of ecotourists to sub-Saharan protected areas. Scientific Reports, 7, 1–9. doi: https://doi.org/10.1038/s41598-017-00858-6. Home, R., Keller, C., Nagel, P., Bauer, N., & Hunziker, M. (2009). Selection criteria for flagship species by conservation organizations. Environmental Conservation, 36(2), 139–148. doi: 10.1017/S0376892909990051. Hoogendoorn, G., & Fitchett, J. M. (2018). Tourism and climate change: a review of threats and adaptation strategies for Africa. Current Issues in Tourism, 21(7), 742–759. https://doi.org/10.1080/13683500.2016.1188893. Hosni, E. (2000). Strategy for sustainable tourism development in the Sahara. UNESCO, Paris, France. Kaján, E., & Saarinen, J. (2013). Tourism, climate change and adaptation: A review. Current Issues in Tourism, 16(2), 167–195. https://doi.org/10.1080/13683500.2013.774323. Komossa, F., van der Zanden, E.H., Schulp, C.J.E., Verburg, P.H. (2018). Mapping landscape potential for outdoor recreation using different archetypical recreation user groups in the European Union. Ecological Indicators, 85, 105-116. doi: 10.1016/j.ecolind.2017.10.015. Kotsiantis, S. B. (2007). Supervised machine learning: A review of classification techniques. Informatica, 31, 249-268. Accessed on 16 August 2020 at http://www.informatica.si/index.php/informatica.

280 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Lanouar, C., Goaied, M. (2019). Tourism, terrorism and political violence in Tunisia: Evidence from Markov-switching models. Tourism Management, 70, 404–418. doi: 10.1016/j.tourman.2018.09.002. Lindsey, P. A., Alexander, R. R., du Toit, J. T., & Mills, M. G. L. (2005). The potential contribution of ecotourism to African wild dog Lycaon pictus conservation in South Africa. Biological Conservation, 123(3), 339–348. https://doi.org/10.1016/j.biocon.2004.12.002. Loarie, S. R., Duffy, P. B., Hamilton, H., Asner, G. P., Field, C. B. & Ackerly, D. D. (2009) The velocity of climate change, Nature, 462: 1052-1055. https://doi.org/10.1038/nature08649 Lu, N., Wang, M., Ning, B., Yu, D., Fu, B. (2018). Research advances in ecosystem services in drylands under global environmental changes. Current Opinion in Environmental Sustainability, 33, 92–98. doi: 10.1016/j.cosust.2018.05.004. Mamadou, M. I., Harouna, A., Habou, M. K. A., Petretto, M., Mahamane, A. (2020). Ethnozoological Knowledge and Local Perceptions about the Reintroduction of the North African Ostrich (Struthio camelus camelus Linnaeus, 1858) in the Koutous, Niger. Journal of Applied Life Sciences International, 23(6), 1-14. Accessed on 16 August 2020 at https://doi.org/10.9734/jalsi/2020/v23i630165. Maron, M. Simmonds, J. S., Watson, J. E. M., Sonter, L. J., Bennun, L., griffiths, V. F., Quétier, F., von Hase, A., Edwards, S., Rainey, H. Bull, J. W., Savy, C. E., Victurine, R., Kiesecker, J., Puydarrieux, P., Stevens, T., Cozannet, N., Jones, J. P. G. (2019). Global no net loss of natural ecosystems. Nature Ecology Evolution, 4, 46–49. doi: https://doi.org/10.1038/s41559-019- 1067-z Marshall, N. A., Marshall, P. A., Abdulla, A., Rouphael, T., & Ali, A. (2011). Preparing for climate change: Recognising its early impacts through the perceptions of dive tourists and dive operators in the Egyptian red sea. Current Issues in Tourism, 14(6), 507–518. https://doi.org/10.1080/13683500.2010.512075. Martinez-Harms, M.J., Bryan, B.A., Balvanera, P., Law, E.A., Rhodes, J.R., Possingham, H.P., Wilson, K.A. (2015). Making decisions for managing ecosystem services. Biological Conservation, 184, 229–238. doi: 10.1016/j.biocon.2015.01.024. Martínez-Harms, M.J., Balvanera, P. (2012). Methods for mapping ecosystem service supply: a review. International Journal of Biodiversity Science, Ecosystem Services & Management, 8(1- 2): 17–25. doi: 10.1080/21513732.2012.663792. McGowan, J., Beaumont, L.J., Smith, R.J. et al. (2020). Conservation prioritization can resolve the flagship species conundrum. Nature Communications, 11, 994. https://doi.org/10.1038/s41467-020-14554-z. FCUP 281 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Milcu, A. I., Hanspach, J., Abson, D., & Fischer, J. (2013). Cultural ecosystem services: A literature review and prospects for future research. Ecology & Society, 18(3), 44–88. https://doi.org/10.5751/ES-05790-180344. Morrison, C., Castley, J.G., Buckley, J.C. (2012) Tourism and the conservation of critically endangered frogs. Plos ONE, 7(9), e43757. doi: 10.1371/journal.pone.0043757. Mossaz, A. C., Buckley, R., & Castley, J. G. (2015). Ecotourism Contributions to Conservation of African Big Cats. Journal for Nature Conservation, 28, 112–118. https://doi.org/10.1016/j.jnc.2015.09.009 Naidoo, R., Fisher, B. Manica, A., Balmford, A. (2016) Estimating economic losses to tourism in Africa from the illegal killing of elephants. Nature Communications, 7, 13379. doi: 10.1038/ncomms13379. NGA (2016). National Geospatial-Intelligence Agency GEOnet Names Server (GNS). Accessed on 15 November of 2018 at http://geonames.nga.mil/gns/html. NIMA (1997). Vector Map (VMap) Level 0. National Imagery and Mapping Agency's (NIMA). Accessed on 06 September 2018 at http://earth-info.nga.mil/publications/vmap0.html. OECD-SWAC (2014). An Atlas of the Sahara-Sahel. Geography, Economics and Security. OECD Publishing. Accessed on 18 June 2018 at http://dx.doi.org/10.1787/9789264222359-en. O’Farrell, P.J., Reyers, B., Le Maitre, D.C., Milton, S.J., Egoh, B., Maherry, A., Colvin, C., Atkinson, D., De Lange, W., Blignaut, J.N., Cowling, R.M. (2010). Multi-functional landscapes in semi arid environments: implications for biodiversity and ecosystem services. Landscape Ecology, 25(8), 1231–1246. doi: 10.1007/s10980-010-9495-9. Oteros-Rozas, E., Martín-López, B., Fagerholm, N., Bieling, C., Plieninger, T. (2018). Using social media photos to explore the relation between cultural ecosystem services and landscape features across five European sites. Ecological Indicators, 94(2), 74-86. https://doi.org/10.1016/j.ecolind.2017.02.009. Paracchini, M. L., Zulian, G., Kopperoinen, L., Maes, J., Schägner, J. P., Termansen, M., Zandersen, M., Perez-Sobad, M., Scholefield, P.A., Bidoglio, G. (2014). Mapping cultural ecosystem services: A framework to assess the potential for outdoor recreation across the EU. Ecological Indicators, 45, 371–385. doi: https://doi.org/10.1016/j.ecolind.2014.04.018.2. Pouzols, F. M., Toivonen, T., Di Minin, E., Kukkala, A. S., Kullberg, P., Kuusterä, J., Lehtomäki, J., Tenkanen, H., Verburg, P. H., Moilanen A. (2014). Global protected area expansion is compromised by projected land-use and parochialism. Nature, 516, 383–386. doi: https://doi.org/10.1038/nature14032.

282 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Rappaport, D. I., Tambosi, L. R., Metzger, J. P. (2015). A landscape triage approach: combining spatial and temporal dynamics to prioritize resoration and conservation. Journal of Applied Ecology, 52: 590-61. doi: https://doi.org/10.1111/1365-2664.12405. Riahi, K., van Vuuren, D. P., Kriegler, E., Edmonds, J., O’Neill, B. C., Fujimori, S., Bauer, N., Calvin, K., Dellink, R., Fricko, O., Lutz, W., Popp, A., Cuaresma, J. C., KC, S., Leimbach, M., Jiang, L., Kram, T., Rao, S., Emmerling, J., … Tavoni, M. (2017). The Shared Socioeconomic Pathways and their energy, land use, and greenhouse gas emissions implications: An overview. Global Environmental Change, 42, 153–168. https://doi.org/10.1016/j.gloenvcha.2016.05.009. Rosa, I.M.D., Pereira, H.M., Ferrier, S., Alkemade, R., Acosta, L.A., Akcakaya, H.R., den Belder, E., Fazel, H.M., Fujimori, S., Harfoot, M., Harhash, K.A., Harrison, P.A., Hauck, J., Hendriks, R.J.J., Hernández, G., Jetz, W., Karlsson-Vinkhuyzen, S.I., Kim, H., King, N., Kok, M.T.J., Kolomytsev, G.O., Lazarova, T., Leadley, P., Lundquist, C.J., Márquez, J.G., Meyer, C., Navarro, L.M., Nesshöver, C., Ngo, H.T., Ninan, K.N., Palomo, M.G., Pereira, L.M., Peterson, G.D., Pichs, R., Popp, A., Purvis, A., Ravera, F., Rondinini, C., Sathyapalan, J., Schipper, A.M., Seppelt, R., Settele, J., Sitas, N., van Vuuren, D. (2017). Multiscale scenarios for nature futures. Nature Ecology & Evolution, 1, 1416-1419. doi: https://doi.org/10.1038/s41559-017- 0273-9. Rosselló, J., Becken, S., & Santana-Gallego, M. (2020). The effects of natural disasters on international tourism: A global analysis. Tourism Management, 79, 1-10. https://doi.org/10.1016/j.tourman.2020.104080. Saarinen, J., Hambira, W. L., Atlhopheng, J., & Manwa, H. (2012). Tourism industry reaction to climate change in Kgalagadi South District, Botswana. Development Southern Africa, 29(2), 273–285. https://doi.org/10.1080/0376835X.2012.675697. Santarém, F., Brito, J.C., Pleguezuelos, J.M., Marques, J., Sousa, F.V., Peyre, O., Geniez, P., Crochet, P.-A. (2018). Birds in northern Mauritania, including range extensions and breeding evidence. Malimbus, 40(2): 37-46. Accessed on 02 August at http://malimbus.free.fr. Santarém, F., & Paiva, F. (2015). Conserving desert biodiversity through ecotourism. Tourism Management Perspectives 16, 176-178. doi: https://doi.org/10.1016/j.tmp.2015.07.016. Santarém, F., Saarinen, J., & Brito, J. C. (2020a). Desert Conservation and Management: Ecotourism. In M. I. Goldstein & D. A. B. T.-E. of the W. B. DellaSala (Eds.), Encyclopedia of the World’s Biomes (pp. 259–273). Elsevier. https://doi.org/https://doi.org/10.1016/B978- 0-12-409548-9.11827-5. Siddig, A. A. H. (2014). Biodiversity of Sudan: between the harsh conditions, political instability FCUP 283 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

and civil wars. Biodiversity Journal, 5(4), 545-555. Accessed on 28 August 2018 at http://www.biodiversityjournal.com. Smith, R. J., Veríssimo, D., Isaac, N. J. B., & Jones, K. E. (2012). Identifying Cinderella species: uncovering mammals with conservation flagship appeal. Conservation Letters, 5(3), 205– 212. doi: 10.1111/j.1755-263X.2012.00229.x. Sow, A. S., Gonçalves, D.V., Sousa, F.V., Martínez-Freiría, F., Santarém, F., Velo-Antón, G., Dieng, H., Campos, J.C., Diagne, S.K., Boratynski, Z., Brito, J.C. (2017). Atlas of the distribution of amphibians and reptiles in the Diawling National Park, Mauritania. Basic and Applied Herpetology, 31, 101-116. doi: http://dx.doi.org/10.11160/bah..87. Steven, R, Castley, J.G., Buckley, R. (2013) Tourism revenue as a conservation tool for threatened birds in protected areas. PLoS ONE, 8(5), e62598. doi: 10.1371/journal.pone.0062598. Strand, J., Soares-Filho, B., Costa, M. H., Oliveira, U., Ribeiro, S. C., Pires, G. F., Oliveira, A., Rajão, R., May, P., van der Hoff, R., Sikamäki, J., da Motta, R. S., Toman, M. (2018). Spatially explicit valuation of the Brazilian Amazon Forest ’ s Ecosystem Services. Nature Sustainability, 1, 657–664. doi: https://doi.org/10.1038/s41893-018-0175-0. Takahashi, Y., Veríssimo, D., MacMillan, D. C., & Godbole, A. (2012). Stakeholder perceptions of potential flagship species for the Sacred Groves of the North Western Ghats, India. Human Dimensions of Wildlife, 17(4), 257–269. doi: 10.1080/10871209.2012.675622. Taylor, N. T., Davis, K. M., Abad, H., McClung, M. R. & Moran, M. D. (2017). Ecosystem services of the Big Bend region of the Chihuahuan Desert. Ecosystem Services, 27, 48–57. doi: https://doi.org/10.1016/j.ecoser.2017.07.017. Telleria, J. L., Ghaillani, H. E. M., Fernandez-Palacios, J. M., Bartolome, J., & Montiano, E. (2008). Crocodiles Crocodylus niloticus as a focal species for conserving water resources in Mauritanian Sahara. Oryx, 42(2), 292–295. https://doi.org/10.1017/s0030605308007850. Tilman, D., Clark, M., Williams, D. R., Kimmel, K., Polasky, S., Packer, C. (2017). Future threats to biodiversity and pathways to their prevention. Nature, 546, 73–81. doi: https://doi.org/10.1038/nature22900. United Nations Educational, Scientific, and Cultural Organization (UNESCO). (2003). Towards a strategy for the sustainable development of tourism in the Sahara, in the context of combating poverty. The Sahara of Cultures and People. Paris, France. United Nations Environmental Programme (UNEP). (2006a). Tourism and deserts: A practical guide to managing the social and environmental impacts in the desert recreation sector (eds.

284 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

de Assis, H. R., & Manca M.). Accessed on 15 January 2018 at http://wedocs.unep.org/handle/20.500.11822/2229. United Nations Environmental Programme (UNEP). (2006b). Global deserts outlook (ed. E. Ezcurra). Accessed on 11 January 2018 at http://wedocs.unep.org/handle/20.500.11822/9581. Vale, C. G., & Brito, J. C. (2015). Desert-adapted species are vulnerable to climate change: Insights from the warmest region on Earth. Global Ecology and Conservation, 4, 369–379. https://doi.org/10.1016/j.gecco.2015.07.012 Vale, C. G., Pimm, S. L., & Brito, J. C. (2015). Overlooked mountain rock pools in deserts are critical local hotspots of biodiversity. PLoS ONE, 10(2), e0118367. https://doi.org/10.1371/journal.pone.0118367. Vale, C. G., Tarroso, P., Brito, J.C. (2015) Predicting species distribution at range margins: testing the effects of study area extent, resolution and threshold selection in the Sahara–Sahel transition zone. Diversity and Distributions, 20, 20-33. https://doi.org/10.1111/ddi.12115. Vaz, A.S., Gonçalves, J.F., Pereira, P., Santarém, F., Vicente, J.R., Honrado, J.P. (2019). Earth observation and social media: Evaluating the spatiotemporal contribution of non-native trees to cultural ecosystem services. Remote Sensing of Environment, 230, 111193. doi: 10.1016/j.rse.2019.05.012. Veríssimo, D., Fraser, I., Girão, W., Campos, A. A., Smith, R. J., Macmillan, D. C. (2014). Evaluating conservation flagships and flagship fleets. Conservation Letters, 7(3), 263–270. doi: 10.1111/conl.12070. Veríssimo, D., Fraser, I., Groombridge, J., Bristol, R., Macmillan, D. C. (2009). Birds as tourism flagship species: a case study of tropical islands Research methods. Animal Conservation, 12(6), 549–558. doi: 10.1111/j.1469-1795.2009.00282.x. Veríssimo, D., MacMillan, D. C., Smith, R. J. (2011). Toward a systematic approach for identifying conservation flagships. Conservation Letters, 4, 1–8. doi: 10.1111/j.1755- 263X.2010.00151.x. Veríssimo, D., Pongiluppi, T., Santos, M. C. M., Develey, P. F., Fraser, I., Smith, R., J., Macmilan, D. C. (2013). Using a systematic approach to select flagship species for bird conservation. Conservation Biology, 28, 269–277. doi: 10.1111/cobi.12142. Veríssimo, D., Vaughan, G., Ridout, M., Waterman, C., Macmillan, D., Smith, R. J. (2017). Increased conservation marketing effort has major fundraising benefits for even the least popular species. Biological Conservation, 211, 95–101. doi: 10.1016/j.biocon.2017.04.018. FCUP 285 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Waldron, A., Mooers, A. O., Miller, D. C., Nibbelink, N., Redding, D., Tyler, S. K., Timmons Roberts, J., &, Gittleman, J. L. (2013). Targeting global conservation funding to limit immediate biodiversity declines. PNAS, 110(29), 12144-12148. doi: https://doi.org/10.1073/pnas.1221370110. Weaver, D. B. (2005). Comprehensive and minimalist dimensions of ecotourism. Annals of Tourism Research, 32(2), 439–455. Doi: https://doi.org/10.1016/j.annals.2004.08.003. Weaver, D. B. (2001). The encyclopedia of ecotourism. CABI, Wallingford, UK. Weiss, C. (2016). Al Qaeda has launched more than 100 attacks in West Africa in 2016. The Long War Journal. Accessed on 23 September 2018 at http://www.longwarjournal.org/archives/2016/06/over-100-al-qaeda-attacks-in-west-africa- since-beginning-of-the-year.php. Willemen, L., Barger, N. N., ten Brink, B., Cantele, M., Erasmus, B. F. N., Fisher, J. L., Gardner, T., Holland, T. G., Kohler, F., Kotiaho, J. S., von Maltitz, G. P., Nangendo, G., Pandit, R., Parrotta, J. A., Potts, M. D., Prince, S. D., Sankaran, M., Brainich, A., Montanarella, L., Scholes, R. (2020). How to halt the global decline of lands. Nature Sustainability, 3, 164–166. doi: https://doi.org/10.1038/s41893-020-0477-x. Wolff, S., Schulp, C.J.E., Verburg, P.H. (2015). Mapping ecosystem services demand: A review of current research and future perspectives. Ecological Indicators, 55, 159-171. doi: https://doi.org/10.1016/j.ecolind.2015.03.016. World Economic Forum (WEF). (2017). The Travel & Tourism Competitiveness Report 2017. Acessed on 18 September 2018 at https://www.weforum.org/reports/the-travel-tourism- competitiveness-report-2017. van Zanten, B.T., Van Berkel, D.B., Meentemeyer, R.K., Smith, J.W., Tieskens, K.F., Verburg, P.H. (2016). Continental-scale quantification of landscape values using social media data. PNAS, 113(46), 12974. doi: www.pnas.org/cgi/doi/10.1073/pnas.1614158113.

286 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Appendices

FCUP 287 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Appendix A

SUPPLEMENTARY INFORMATION FOR ARTICLE II

288 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

New method to identify and map flagship fleets for promoting conservation and ecotourism

Frederico Santarém1,2,3, Paulo Pereira1,2, Jarkko Saarinen3,4, José Carlos Brito1,2

1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 – Geography Research Unit, University of Oulu, Finland 4 – School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

List of Supplementary Texts, Figures and Tables:

Text A1 – Sources used to collect species data. Text A2 – Assessing the contribution of variables that explain species’ flagship appealing and grouping species according to their shared characteristics to flagship suitability Text A3 – Explanation and results of the Bayesian Information Criterion (BIC) models analysed to identify the most likely number of clusters. Fig. A2 – Hypothetical representation of animal’ body patterns considered in this study. Table A1 – The nine flagship fleets identified by the Bayesian Information Criterion (BIC). FCUP 289 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Text A1 – Sources used to collect species data

We accessed species traits from several data sources. For amphibians we used Schleich, Kästle and Kabisch (1996), Rödel (2000), Largen and Spawls (2010), and AmphibiaWeb (2017). For reptiles we used Din (2006), Trape and Mané (2006), Sindaco and Jeremcenko (2008), Trape et al. (2012), and Sindaco et al. (2013). For mammals we used Kingdon (1997), Pacifici et al. (2013), and Bennie et al. (2014). For birds we used Thomas and Robin (1977), Maclean (1983), del Hoyo et al. (1992), Borrow and Demey (2001), Madge et al. (2002), Jonsson (2006), Sinclair and Ryan (2010), and HBW (2016). We complemented the data with generalist references (Jones et al., 2009; Myhrvold et al., 2015; and Encyclopedia of Life, 2017) and expert knowledge whenever needed. For those species that we were unable to find traits data in the literature, we used the median for the or likely species (when that species had no other representatives from the same genus) as in Smith et al. (2012) and Veríssimo et al. (2017).

References

AmphibiaWeb, 2017. AmphibiaWeb. Available at https://amphibiaweb.org/, Accessed date: 3 March 2017. Bennie, J.J., Duffy, J.P., Inger, R., Gaston, K.J., 2014. Biogeography of time partitioning in mammals. Proc. Natl. Acad. Sci. 111, 13727–13732. https://doi.org/10.1073/pnas.1216063110. Borrow, N., Demey, R., 2001. Birds of Western Africa. Cristopher Helm, London, UK. del Hoyo, J., Elliott, A., and Sargatal, J., 1992. Handbook of the Birds of the World: Ostrich to ducks. Vol. 1. Lynx Editions. Din, S.B., 2006. A Guide to Reptiles and Amphibians of Egypt. American University in Cairo Press, Cairo, Egypt. HBW, 2016. Handbook of the Birds of the World. Available at https://www.hbw.com/, Accessed date: 22 December 2016). IUCN, 2017. The IUCN Red List of Threatened Species, 2017-3. International Union for Conservation of Nature and Natural Resources. Available at http://www.iucnredlist.org/, Accessed date: 15 November 2017. Jones, K.E., Bielby, J., Cardillo, M., Fritz, S.A., O’Dell, J., Orme, C.D.L., … Purvis, A., 2009. PanTHERIA: A species-level database of life history, ecology, and geography of extant and recently extinct mammals. Ecology, 90, 2648. https://doi.org/10.1890/08-1494.1.

290 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Jonsson, L., 2006. Birds of Europe: With North Africa and the Middle East. Princeton University Press, New Jersey, USA. Kingdon, J., 1997. The Kingdon Field Guide to African Mammals. Academic Press, California, USA. Largen, M., and Spawls, S., 2010. The Amphibians and Reptiles of Ethiopia and Eritrea. Frankfurt: Edition Chimaira. Encyclopedia of Life, 2017. Encyclopedia of Life. Available at www.eol.org, Accessed date: 14 August 2017. Maclean, G.L., 1983. Water transport by sandgrouse. BioScience, 33(6), 365–369. https://doi.org/10.2307/1309104. Madge, S., McGowan, P.J.K., Kirwan, G.M., 2002. Pheasants, Partridges and Grouse: A Guide to the Pheasants, Partridges, Quails, Grouse, Guineafowl, Buttonquails and Sandgrouse of the World. Christopher Helm, London, UK. Myhrvold, N.P., Elita, B., Benjamin, C., Sivam, D., Freeman, D.L., Ernest, S.K.M., 2015. An amniote life-history database to perform comparative analyses with birds, mammals, and reptiles. Ecology, 96(11), 3109. https://doi.org/10.1890/15-0846R.1. Pacifici, M., Santini, L., Marco, M. Di, Baisero, D., Francucci, L., Marasini, G.G., … Rondinini, C., 2013. Generation length for mammals. Nature Conservation, 5, 87–94. 10.3897/natureconservation.5.5734. Rödel, M.O., 2000. Herpetofauna of West Africa: Amphibians of the West African Savanna. Edition Chimaira, Frankfurt, Germany. Schleich, H.H., Kästle, W., Kabisch, K., 1996. Amphibians and Reptiles of North Africa: Biology, Systematics, Field Guide. Koeltz Scientific Books, Königstein, Germany. Sinclair, I., Ryan, P., 2010. Birds of Africa, South of the Sahara. Struik Nature, Cape Town, South Africa. Sindaco, R., Jeremcenko, V.K., 2008. The Reptiles of the Western Palearctic, Volume 1: Annotated Checklist and Distributional Atlas of the , Crocodiles, Amphisbaenians and Lizards of Europe, North Africa, Middle East and Central Asia. Edizioni Belvedere, Latina, Italy. Sindaco, R., Jeremčenko, V.K., Venchi, A., Grieco, C., 2013. The Reptiles of the Western Palearctic: Annotated Checklist and Distributional Atlas of the Snakes of Europe, North Africa, Middle East and Central Asia, with an update to the vol. 1. Edizioni Belvedere, Latina, Italy. Smith, R.J., Veríssimo, D., Isaac, N.J.B., Jones, K.E., 2012. Identifying Cinderella species: Uncovering mammals with conservation flagship appeal. Conservation Letters, 5, 205–212. https://doi.org/10.1080/10871200701883408. FCUP 291 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Thomas, D. H., and Robin, P., 1977. Comparative studies of thermoregulatory and osmoregulatory behaviour and physiology of five species of sandgrouse (Aves: Pterocliidae) in Morocco. Journal of , 183, 229–249. https://doi.org/10.1111/j.1469- 7998.1977.tb04184.x. Trape, J.-F., and Mané, Y., 2006. Guide des Serpents d’Afrique Occidentale: Savane et désert. IRD Édition, Paris, France. Trape, J.-F., Trape, S., Chirio, L., 2012. Lézards, Crocodiles et Tortues d’Afrique Occidentale et du Sahara. IRD Éditions, Marseille, France. United Nations Development Programme (UNDP) (2019). Human Development Report 2019: Beyond Income, Beyond Averages, Beyond Today - Inequalities in Human Development in the 21st Century. Accessed on 17 February 2020 at https://www.un-ilibrary.org/economic-and- social-development/human-development-report-2019_838f78fd-en. United Nations World Tourism Organization (UNWTO) (2019). Yearbook of Tourism Statistics, Data 2013-2017, 2019 Edition. Acessed on 23 February 2020 at https://www.e- unwto.org/doi/book/10.18111/9789284420414. Veríssimo, D., Vaughan, G., Ridout, M., Waterman, C., Macmillan, D., Smith, R.J., 2017. Increased conservation marketing effort has major fundraising benefits for even the least popular species. Biological Conservation, 211, 95–101. https://doi.org/10.1016/j.biocon.2017.04.018.

292 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Text A2 – Assessing the contribution of variables that explain species’ flagship appealing and grouping species according to their shared characteristics to flagship suitability

To determine the contribution of each species trait to the flagship appealing character, we performed a Principal Component Analysis (PCA) with metrics, which allows the combination of numerical and categorical data. To perform the PCA with mixed data, we first scaled the numerical data by subtracting each variable value by its mean and then dividing by its standard deviation. This process converts each raw datum score into a standardized value with mean 0 and standard deviation 1. We performed a mixed PCA using the R package PCAmixdata (Chavent et al., 2017a), dedicated to the analysis of multivariate data where the observations are described by a mixture of numerical and categorical variables. A PCA with metrics is used within the package, which is a generalization of the standard PCA where metrics are used to add weights on the observations and on the variables of the multivariate data matrix. The framework performs a standard PCA for numerical data and a standard Multiple Correspondence Analysis (MCA) for categorical data. In the standard PCA, the n observations are weighted by 1 and the p columns are weighted by 1, 푛 being the distance between two observations the Euclidean distance between two rows of the original matrix for numerical data. Hence, each loading is the linear correlation between the numerical variable and the principal component, whereas each eigenvalue is the variance of each principal component. In a standard MCA, each categorical variable m has j levels 푗 and ∑푖=1 푚푖 = 푚. Knowing that the factor coordinates of the observations (rows in a data matrix) and the factor coordinates of the levels (columns in a data matrix) are obtained by applying a PCA on two different matrices – the matrix of the row profiles and the matrix of the column profiles, the 1 푛 n observations (rows) are weighted by and the m levels (columns) are weighted by , where 푛푠 푛 푛푠 denotes the number of observations that belong to the sth level. The distance between two observations is then a weighted Euclidean distance similar to the χ2 distance in a Correspondence Analysis, which gives more importance to rare levels. Hence, each coordinate (similar to the loading of the standard PCA), is the mean value of the standardized factor coordinates of the observations that belong to the level s, whereas each eigenvalue is the sum of the correlation ratios between the p categorical variables and the ith principal component (Chavent et al., 2017b). The PCA with metrics is based on a Generalized Singular Value Decomposition (GSVD) of pre- processed data. The first step, the pre-processing of data, builds the real matrix Z that combines the standardized version of the numerical data matrix and the centred version of the indicator FCUP 293 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel matrix of the categorical data matrix. It then builds the diagonal matrix N of the weights of the rows of Z, and the diagonal matrix M of the weights of the columns of Z. The second step, the factor coordinates processing, builds a matrix A of dimension (p1+m)r containing the factor coordinates of the p1 numerical variables and the factor coordinates of the m levels of the p2 categorical variables. Finally, in the third step, the squared loadings – defined as the contributions of the variables to the variance of the principal components – are processed. The contributions of the variables are calculated directly from the matrix A. One obtains the squared Euclidean distance on the standardized numerical variables plus the squared χ2 distance calculated on the levels of the categorical variables. Note that the contribution of a categorical variable is the sum of the contributions of its levels (Chavent et al., 2017b). When the original data set is mixed, the barycentric property in MCA remains true, i.e., the coordinates of the categories are the averages of the standardized component scores of the objects in those categories (Chavent et al. 2012). The contribution of each species trait for the appealing character of the species is given by the link between the variables and the principal components. A list of the squared loadings of the multivariate data shows these contributions. Squared loadings are the squared correlation

2 2 r (fi,xj) for numerical variables and the correlation ratios ŋ (fi|xj) for categorical variables, respectively (Chavent et al., 2012; Chavent et al., 2017b). To simplify the interpretation of the principal components, we applied a rotation to the loadings. The idea was to obtain either large (close to 1) or small (close to 0) loadings, in order to associate variables with the principal components. The function implements a generalization of the varimax procedure to mixed (numerical and categorical) data, originally proposed by Kaiser (Kaiser, 1958), and is applied to the first q principal components of the previous procedure. Kaiser introduced the varimax criterion aiming at maximizing the sum over the columns of the squared elements of the loading matrix, which contains the correlations between the variables and the principal components (Kaiser, 1958). This criterion was then extended to the principal component method mixing numerical and categorical data. The varimax criterion is then expressed with squared loadings defined as correlation ratios for categorical variables and squared correlations for numerical variables (Kaiser, 1991). At each iteration, a planar rotation is applied, for which the rotation matrix T only depends on an angle Ɵ. This procedure rotates pairs of dimensions (1, 2, …, k) in the following way: dimensions 1 and 2, 1 and 3, …, 1 and k, 2 and 3, …, k -1 and k, iteratively until the processes converge, i.e., until an angle of rotation equal to zero is obtained (Chavent et al., 2012). For numerical data, PCArot is the standard varimax procedure defined by Kaiser (1958) for rotation in PCA. For categorical data, PCArot is an orthogonal rotation procedure for MCA. The interpretation rules associated with the barycentric property remains true after

294 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

rotation. Hence, the squared loadings after rotation are given by the squared correlations between the numerical (and by the correlation ratios between the categorical) variables and the rotated components (Chavent et al., 2012). Model-based clustering is a popular method for unsupervised learning when prior knowledge of the number of groups or any information regarding their composition is unknown. It relies on a probabilistic assumption about data distribution (contrary to classic heuristic method such as k-means), meaning that approximate Bayes factors can be used to compare models with different parametrizations. Parameters of the Gaussian mixture models are estimated by an iterative expectation-maximization algorithm (Fraley and Raftery, 1998). The Bayesian Information Criterion (BIC) tests different hypotheses to select the number of mixture components (i.e. the number of clusters). The information criterion is based on maximized log-likelihood where a penalty is added, based on the number of model parameters. As the likelihood increases with the addition of clusters, a penalty term for the number of estimated parameters is subtracted from the log-likelihood. The larger the value of BIC, the stronger the evidence for a given model and number of groups (Fraley and Raftery, 1998). Mclust is a popular R package for model-based clustering based on finite Gaussian mixture modelling (Scrucca et al., 2016; Fraley et al., 2017). We first applied a dimension-reduction technique, where the two first rotated principal components of the PCA were used as input for the clustering analysis. We then tested fourteen models with different volume, shape, and orientation to evaluate the number of flagship groups (see Text A3).

References

Chavent, M., Kuentz-Simonet, V., Labenne, A., Liquet, B., Saracco, J., 2017a. PCAmixdata: Multivariate Analysis of Mixed Data. R package version 3.1. Available at https://cran.r- project.org/package=PCAmixdata, Accesed date: 13 March 2018. Chavent, M., Kuentz-Simonet, V., Labenne, A., Saracco, J., 2017b. Multivariate Analysis of Mixed Data: The R Package PCAmixdata. Chavent, M., Kuentz-Simonet, V., Saracco, J., 2012. Orthogonal rotation in PCAMIX. Advances in Data Analysis and Classification, 6, 131–146. https://doi.org/10.1007/s11634-012-0105-3. Fraley, C., Raftery, A.E., 1998. How many clusters? Which clustering method? Answers via model-based cluster analysis. The Computer Journal, 41, 578–588. https://doi.org/10.1093/comjnl/41.8.578. Fraley, C., Raftery, A.E., Scrucca, L., Murphy, T.B., and Fop, M., 2017. Package “mclust”. R FCUP 295 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

package version 5.4. Kaiser, H.F., 1958. The varimax criterion for analytic rotation in factor analysis. Psychometrika, 23(3), 187–200. https://doi.org/10.1007/BF02289233. Kaiser, H.F., 1991. Simple structure in component analysis techniques for mixtures of qualitative and quantitative variables. Psychometrika, 56, 197–212. https://doi.org/10.1007/BF02294458 Scrucca, L., Fop, M., Murphy, T.B., Raftery, A.E., 2016. mclust 5: Clustering, classification and density estimation using Gaussian finite mixture models. R J., 8, 289–317.

296 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Text A3 – Explanation and results of the Bayesian Information Criterion (BIC) models analysed to identify the most likely number of clusters.

The BIC partitioned the set of observations into k sets (clusters), minimising within cluster variance. In order to partition the data, several possible models exist. The BIC method here implemented used 14 models, which are the combination of different model types: 1) the distribution of the geometric construct used to split the data (diagonal, spherical or ellipsoidal); 2) the volume of the geometric construct (equal across clusters, unequal across clusters, applies to all three geometric constructs); 3) the shape of the construct (equal shape or varying shape, applies to both spherical and ellipsoidal models); and 4) the orientation (equal or varying orientation). The list of models is defined as follows (obtained from Scrucca et al., 2016): "EII" = spherical, equal volume "VII" = spherical, unequal volume "EEI" = diagonal, equal volume and shape "VEI" = diagonal, varying volume, equal shape "EVI" = diagonal, equal volume, varying shape "VVI" = diagonal, varying volume and shape "EEE" = ellipsoidal, equal volume, shape, and orientation "EVE" = ellipsoidal, equal volume and orientation "VEE" = ellipsoidal, equal shape and orientation "VVE" = ellipsoidal, equal orientation "EEV" = ellipsoidal, equal volume and equal shape "VEV" = ellipsoidal, equal shape "EVV" = ellipsoidal, equal volume "VVV" = ellipsoidal, varying volume, shape, and orientation

For each possible number of clusters defined and for each model, the BIC models identify the centroids that best partition the data and compute the log-likelihood of the solution by comparing the k clusters with the k-1 solution and determine if k clusters performed better than k- 1. BIC tends to unambiguously prefer solutions with less clusters and more dissociated from each other. After that solution is found the model is again projected back onto the data. These models can be thought of a group of k polygons that one can use to separate the observations (species) into k groups (flagship fleets). FCUP 297 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

In the figure, relationships between BIC and number of components are depicted for fourteen multiple multivariate models. Codes on the small box on the right bottom of the figure are parametrizations of the covariance matrix of the multidimensional data in the Gaussian models for hierarchical clustering, which vary according to geometric (distribution, volume, shape, and orientation) distribution of the data. The model with the highest BIC value is the ellipsoidal, varying volume, shape, and orientation (VVV) with nine components. The values of this model increase with the number of clusters and most of the increase occurs with less than nine components, which were chosen as the optimum number of clusters defining flagship fleets. Higher number of clusters brings additional complexity to the analysis without a substantial increment of the BIC value (Scrucca et al., 2016).

Reference Scrucca, L., Fop, M., Murphy, T.B., Raftery, A.E., 2016. mclust 5: Clustering, classification and density estimation using Gaussian finite mixture models. R J., 8, 289–317.

298 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. A1. Sahara-Sahel limits. Sahara-Sahel limits in Africa (small inset) following Dinerstein et al. (2017), range countries within the area, and main mountains and waterbodies (numbered clockwise). BF – Burkina Faso; CAM – Cameroon; ER – Eritrea; TU - Tunisia.

Reference Dinerstein, E., Olson, D., Joshi, A., Vynne, C., Burgess, N.D., Wikramanayake, E., … Kindt, R., 2017. An ecoregion-based approach to protecting half the terrestrial realm. BioScience, 67, 534– 545. https://doi.org/10.1093/biosci/bix014.

FCUP 299 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. A2. Hypothetical representation of animal’ body patterns considered in this study.

300 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table A1. The nine flagship fleets identified by the Bayesian Information Criterion (BIC). Description, class and species (Taxa) – alongside their endemism (END; Sahara or Sahel) and IUCN conservation (CS) status – that belong to each of the fleets. Species data follow IUCN (2017). Fleet Description Class Taxa END CS A Large-sized and heavy mammals and reptiles, approximately half of them are regional endemics exhibiting some unique desert adaptations, more than a half is threatened with extinction, and most are herbivorous Mammalia Addax nasomaculatus Sahara CR Ammotragus lervia Sahara VU

Equus africanus CR

Eudorcas rufifrons Sahel VU

Gazella cuvieri Sahara EN

Gazella dorcas Sahara VU

Gazella leptoceros Sahara EN

Giraffa camelopardalis LC

Hippopotamus amphibius VU

Kobus megaceros EN

Loxodonta Africana VU

Nanger dama Sahara CR

Nanger soemmerringii VU

Oryx beisa NT

Oryx dammah Sahara EW

Panthera leo VU

Syncerus caffer LC

Trichechus senegalensis VU

Reptilia Acanthodactylus spinicauda Sahara CR

Crocodylus niloticus LC

Crocodylus suchus NE

Naja nubiae Sahel NE

Pseudocerastes fieldi LC

Pseudocerastes persicus LC

Python sebae NE

Testudo kleinmanni Sahara CR

Uromastyx acanthinura Sahara NE FCUP 301 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Uromastyx aegyptia VU

Uromastyx alfredschmidti Sahara NT

Uromastyx dispar Sahara NE

Uromastyx geyri Sahara NE

Uromastyx nigriventris Sahara NE

Uromastyx occidentalis Sahara NE

Uromastyx ocellata LC

Uromastyx ornata LC

Varanus griseus NE B Small-sized and light birds, mammals and reptiles, most are regional endemics, exhibit several unique desert adaptations, are not threatened, and are carnivorous Aves Passer cordofanicus Sahel LC Passer luteus Sahel LC

Pterocles alchata LC

Pterocles coronatus LC

Pterocles exustus LC

Pterocles gutturalis LC

Pterocles lichtensteinii LC

Pterocles orientalis LC

Pterocles quadricinctus LC

Pterocles senegallus LC

Mammalia Crocidura tarfayensis Sahara DD

Ctenodactylus vali Sahara DD

Ictonyx libyca Sahara LC

Jaculus jaculus LC

Jaculus orientalis LC

Vulpes zerda Sahara LC

Reptilia Acanthodactylus aegyptius Sahara NE

Acanthodactylus aureus Sahara NE

Acanthodactylus boskianus NE

Acanthodactylus dumerili Sahara NE

Acanthodactylus longipes Sahara NE

Acanthodactylus opheodurus LC

Acanthodactylus scutellatus NE

302 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Acanthodactylus taghitensis Sahara DD

Agama boueti Sahel LC

Agama boulengeri Sahara LC

Agama spinosa LC

Agama tassiliensis Sahara LC

Cerastes cerastes NE

Cerastes vipera LC

Chalcides boulengeri Sahara NE

Chalcides delislei Sahel LC

Chalcides humilis Sahara NE

Chalcides sepsoides Sahara LC

Chalcides sphenopsiformis Sahara LC

Dasypeltis sahelensis Sahel NE

Echis coloratus NE

Echis pyramidum LC

Eryx jaculus NE

Leptotyphlops algeriensis Sahara NE

Leptotyphlops boueti Sahel NE

Leptotyphlops cairi Sahara NE

Leptotyphlops NE macrorhynchus

Leptotyphlops nursii NE

Mauremys leprosa NE

Mesalina rubropunctata Sahara NE

Philochortus lhotei Sahel NE

Pristurus adrarensis Sahara DD

Pseudotrapelus sinaitus NE

Ptyodactylus guttatus NE

Ptyodactylus hasselquistii NE

Ptyodactylus oudrii LC

Ptyodactylus ragazzi NE

Ptyodactylus siphonorhina Sahara NE

Scincopus fasciatus Sahel DD

Scincus albifasciatus Sahara NE FCUP 303 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Scincus scincus NE

Stellagama stellio LC

Stenodactylus petri Sahara NE

Stenodactylus stenodactylus Sahara NE

Trapelus boehmei Sahara LC

Trapelus mutabilis Sahara NE

Trapelus pallidus NE

Trapelus schmitzi Sahara DD

Trapelus tournevillei Sahara LC

Tropiocolotes algericus Sahara NE

Tropiocolotes bisharicus Sahara NE

Tropiocolotes steudeneri Sahara NE

Typhlops etheridgei Sahara DD

Typhlops vermicularis NE C Medium-sized birds, mammals and reptiles, none is endemic, almost none with unique desert adaptations, most are not threatened with extinction, and almost half are carnivorous or herbivorous

Aves Ardea cinerea LC

Ardea goliath LC

Ardeotis arabs NT

Balaeniceps rex VU

Balearica pavonina VU

Chlamydotis undulata VU

Circaetus beaudouini VU

Cyanochen cyanoptera VU

Dendrocygna bicolor LC

Gypaetus barbatus LC

Gyps africanus EN

Gyps fulvus LC

Gyps rueppellii EN

Necrosyrtes monachus EN

Neophron percnopterus EN

Plectropterus gambensis LC

Polemaetus bellicosus NT

304 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Sagittarius serpentarius VU

Struthio camelus LC

Thalassornis leuconotus LC

Torgos tracheliotos VU

Trigonoceps occipitalis VU

Mammalia Acinonyx jubatus VU

Alcelaphus buselaphus LC

Aonyx capensis LC

Arvicanthis ansorgei LC

Canis aureus LC

Canis mesomelas LC

Capra nubiana VU

Caracal caracal LC

Cervus elaphus LC

Civettictis civetta LC

Colobus guereza LC

Crocuta crocuta LC

Damaliscus lunatus LC

Eidolon helvum NT

Felis chaus LC

Felis silvestris LC

Heterohyrax brucei LC

Hippotragus equinus LC

Hystrix cristata LC

Hystrix indica LC

Kobus ellipsiprymnus LC

Kobus kob LC

Lemniscomys macculus LC

Lemniscomys zebra LC

Leptailurus serval LC

Lepus capensis LC

Lepus habessinicus LC

Lepus microtis LC

Lutra lutra NT FCUP 305 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Lutra maculicollis LC

Lycaon pictus EN

Madoqua saltiana LC

Nesokia indica LC

Oreotragus oreotragus LC

Orycteropus afer LC

Otomys typus LC

Ourebia ourebi LC

Panthera pardus NT

Papio anubis LC

Phacochoerus africanus LC

Potamochoerus larvatus LC

Potamochoerus porcus LC

Procavia capensis LC

Rattus norvegicus LC

Redunca redunca LC

Rhinolophus mehelyi VU

Smutsia temminckii LC

Sus scrofa LC

Tachyoryctes splendens LC

Theropithecus gelada LC

Tragelaphus derbianus LC

Tragelaphus scriptus LC

Tragelaphus spekii LC

Tragelaphus strepsiceros LC

Vormela peregusna VU

Reptilia Acanthodactylus pardalis VU

Centrochelys sulcata VU

Chamaeleo chamaeleon LC

Daboia mauritanica NT

Dendroaspis polylepis LC

Malpolon monspessulanus LC

Platyceps florulentus LC

Platyceps rogersi LC

306 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Python regius LC

Testudo graeca VU

Varanus exanthematicus LC

Walterinnesia aegyptia LC D Small to medium-sized amphibians, birds, mammals and reptiles, the majority is non-endemic, almost none with unique desert adaptations, most of them were never evaluated for conservation status, and more than a half is carnivorous Amphibia Amietophrynus kassasii Sahara LC

Aves Burhinus oedicnemus LC

Neotis nuba Sahara NT

Mammalia Allactaga tetradactyla VU

Desmodilliscus braueri Sahel LC

Felis margarita NT

Felovia vae Sahel DD

Genetta genetta LC

Hyaena hyaena NT

Lemniscomys hoogstraali Sahel DD

Massoutiera mzabi Sahara LC

Psammomys obesus LC

Psammomys vexillaris Sahara DD

Sekeetamys calurus Sahara LC

Taterillus pygargus Sahel LC

Vulpes pallida Sahel LC

Vulpes rueppellii LC

Reptilia Atractaspis micropholis Sahel LC

Atractaspis watsoni NE

Bitis arietans NE

Crotaphopeltis hotamboeia NE

Gongylophis colubrinus NE

Gongylophis muelleri NE

Hemorrhois algirus Sahara LC

Hemorrhois dorri NE

Hemorrhois nummifer NE

Lamprophis fuliginosus NE FCUP 307 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Lythorhynchus diadema NE

Malpolon insignitus NE

Naja haje LC

Naja nigricollis NE

Ophisops elbaensis DD

Pelomedusa subrufa NE

Philochortus zolii Sahara CR

Platyceps saharicus Sahara NE

Psammophis aegyptius Sahara NE

Psammophis elegans NE

Psammophis punctulatus NE

Psammophis schokari NE

Psammophis sibilans NE

Scutophis moilensis NE

Spalerosophis diadema NE

Spalerosophis dolichospilus DD

Tarentola chazaliae Sahara VU

Telescopus obtusus NE

Telescopus tripolitanus NE

Trapelus savignii VU

Trionyx triunguis NE

Varanus niloticus NE E Small-sized amphibians, birds, mammals and reptiles, half are regional endemics, almost none with unique desert adaptations, half of them were never evaluated for conservation status and the other half is not threatened, and the majority is carnivorous Amphibia Kassina wazae Sahel DD Aves Anthoscopus punctifrons Sahel LC

Caprimulgus eximius Sahel LC

Dendropicos elachus Sahel LC

Mirafra cordofanica Sahara LC

Prinia fluviatilis Sahel LC

Spiloptila clamans Sahel LC

Turdoides fulva Sahara LC

308 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Mammalia Acomys airensis Sahel LC

Acomys cahirinus Sahara LC

Acomys seurati Sahara LC

Crocidura lusitania Sahel LC

Crocidura pasha Sahel LC

Eptesicus floweri Sahel LC

Gerbillus bottai Sahel DD

Gerbillus gerbillus Sahara LC

Gerbillus lowei Sahel DD

Gerbillus muriculus Sahel DD

Gerbillus nancillus Sahel DD

Gerbillus principulus Sahel DD

Gerbillus pyramidum Sahel LC

Gerbillus stigmonyx Sahel DD

Gerbillus tarabuli Sahara LC

Grammomys aridulus Sahel DD

Mastomys kollmannspergeri Sahel LC

Pipistrellus deserti Sahara LC

Plecotus christii Sahara DD

Rhinopoma hardwickii LC

Rhinopoma microphyllum LC

Reptilia Acanthodactylus busacki Sahara LC

Agama impalearis LC

Chalcides ocellatus NE

Chamaeleo africanus NE

Eumeces schneideri NE

Hemidactylus angulatus NE

Hemidactylus flaviviridis NE

Hemidactylus robustus NE

Hemidactylus sinaitus NE

Heremites vittatus NE

Latastia longicaudata NE

Mesalina guttulata NE

Mesalina martini NE FCUP 309 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Mesalina olivieri NE

Mesalina pasteuri Sahara LC

Ophisops elegans NE

Pristurus flavipunctatus NE

Pseuderemias mucronata NE

Stenodactylus mauritanicus NE

Tarentola annularis NE

Tarentola boehmei Sahara LC

Tarentola deserti Sahara LC

Tarentola ephippiata NE

Tarentola hoggarensis NE

Tarentola mindiae Sahara LC

Tarentola neglecta Sahara LC

Tarentola parvicarinata NE

Tarentola senegambiae NE

Trachylepis quinquetaeniata NE

Tropiocolotes nattereri LC

Tropiocolotes nubicus Sahara DD

Tropiocolotes tripolitanus Sahara LC F Small-sized nocturnal mammals and reptiles, more than a half are regional endemics, without unique desert adaptations, not threatened, and omnivorous

Amphibia Pseudepidalea variabilis DD

Mammalia Crocidura floweri DD

Crocidura planiceps DD

Crocidura religiosa DD

Gerbillus amoenus Sahara LC

Gerbillus floweri LC

Gerbillus latastei Sahara LC

Gerbillus mackillingini Sahara LC

Gerbillus nigeriae Sahel LC

Gerbillus perpallidus Sahara LC

Gerbillus rosalinda Sahel LC

Gerbillus rupicola Sahel LC

Gerbillus watersi Sahara LC

310 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Pipistrellus ariel DD

Pipistrellus hanaki DD

Steatomys cuppedius Sahel LC

Tadarida ventralis DD

Taterillus arenarius Sahel LC

Taterillus lacustris Sahel LC

Taterillus petteri Sahel LC

Taterillus tranieri Sahel LC

Reptilia Letheobia erythraea DD G Small to medium-sized amphibians, birds, mammals and reptiles, not endemic, without unique desert adaptations, not threatened, and the majority is carnivorous

Amphibia Afrixalus quadrivittatus LC

Afrixalus vittiger LC

Amietia angolensis LC

Amietophrynus asmarae LC

Amietophrynus maculatus LC

Amietophrynus regularis LC

Amietophrynus steindachneri LC

Amietophrynus xeros LC

Bufo bufo LC

Bufo mauritanicus LC

Bufo pentoni LC

Discoglossus pictus LC

Discoglossus scovazzi LC

Hoplobatrachus occipitalis LC

Hyla meridionalis LC

Hylarana galamensis LC

Hyperolius acuticeps LC

Kassina senegalensis LC

Pelophylax bedriagae LC

Pelophylax saharicus LC

Phrynobatrachus francisci LC

Phrynobatrachus latifrons LC

Phrynobatrachus natalensis LC FCUP 311 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Pseudepidalea boulengeri LC

Pseudepidalea brongersmai NT

Ptychadena bibroni LC

Ptychadena mascareniensis LC

Ptychadena pumilio LC

Ptychadena schillukorum LC

Ptychadena tellinii LC

Ptychadena trinodis LC

Pyxicephalus edulis LC

Tomopterna cryptotis LC

Xenopus clivii LC

Xenopus muelleri LC

Aves Anastomus lamelligerus LC

Anhinga rufa LC

Aquila chrysaetos LC

Aquila fasciata LC

Aquila rapax LC

Aquila verreauxii LC

Aquila wahlbergi LC

Ardea melanocephala LC

Ardea purpurea LC

Ardeola ralloides LC

Asio abyssinicus LC

Asio capensis LC

Asio otus LC

Athene noctua LC

Bostrychia carunculata LC

Bostrychia hagedash LC

Bubo africanus LC

Bubo ascalaphus LC

Bubo bubo LC

Bubo lacteus LC

Bubulcus ibis LC

Bucorvus abyssinicus LC

312 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Burhinus capensis LC

Burhinus senegalensis LC

Buteo augur LC

Buteo auguralis LC

Buteo rufinus LC

Butorides striata LC

Bycanistes subcylindricus LC

Caprimulgus aegyptius LC

Caprimulgus climacurus LC

Caprimulgus inornatus LC

Caprimulgus natalensis LC

Caprimulgus poliocephalus LC

Caprimulgus tristigma LC

Casmerodius albus LC

Ceryle rudis LC

Charadrius alexandrinus LC

Chelictinia riocourii LC

Chrysococcyx caprius LC

Ciconia abdimii LC

Ciconia episcopus LC

Circaetus cinerascens LC

Circaetus cinereus LC

Circaetus pectoralis LC

Clamator glandarius LC

Clamator levaillantii LC

Columba LC

Corvus albus LC

Corvus capensis LC

Corvus corax LC

Corvus corone LC

Corvus crassirostris LC

Corvus rhipidurus LC

Corvus ruficollis LC

Cuculus clamosus LC FCUP 313 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Cuculus gularis LC

Cuculus solitarius LC

Cursorius cursor LC

Dendrocygna viduata LC

Dendropicos abyssinicus LC

Egretta ardesiaca LC

Egretta garzetta LC

Ephippiorhynchus LC senegalensis

Falco biarmicus LC

Falco concolor NT

Falco tinnunculus LC

Francolinus albogularis LC

Francolinus bicalcaratus LC

Francolinus clappertoni LC

Francolinus coqui LC

Francolinus erckelii LC

Francolinus icterorhynchus LC

Francolinus leucoscepus LC

Francolinus schlegelii LC

Francolinus squamatus LC

Glaucidium perlatum LC

Gorsachius leuconotus LC

Haliaeetus vocifer LC

Hieraaetus spilogaster LC

Lamprotornis pulcher Sahel LC

Leptoptilos crumeniferus LC

Lophaetus occipitalis LC

Lybius vieilloti LC

Macheiramphus alcinus LC

Macronyx croceus LC

Macronyx flavicollis NT

Melierax metabates LC

Mesophoyx intermedia LC

314 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Milvus migrans LC

Mycteria ibis LC

Neotis denhami NT

Netta erythrophthalma LC

Numida meleagris LC

Oenanthe deserti LC

Otus leucotis LC

Otus scops LC

Otus senegalensis LC

Pelecanus rufescens LC

Phalacrocorax africanus LC

Phalacrocorax carbo LC

Platalea alba LC

Plegadis falcinellus LC

Podica senegalensis LC

Polyboroides typus LC

Rostratula benghalensis LC

Rynchops flavirostris NT

Sarkidiornis melanotos LC

Scopus umbretta LC

Scotopelia peli LC

Sporopipes frontalis LC

Stephanoaetus coronatus NT

Streptopelia turtur LC

Strix aluco LC

Strix butleri LC

Strix woodfordii LC

Terathopius ecaudatus NT

Threskiornis aethiopicus LC

Thripias namaquus LC

Trachyphonus margaritatus LC

Tricholaema melanocephala LC

Tyto alba LC

Upupa epops LC FCUP 315 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Vanellus tectus LC

Mammalia Aethomys stannarius DD

Arvicanthis abyssinicus LC

Arvicanthis niloticus LC

Asellia tridens LC

Atilax paludinosus LC

Atlantoxerus getulus LC

Canis adustus LC

Chlorocebus aethiops LC

Chlorocebus sabaeus LC

Chlorocebus tantalus LC

Cricetomys gambianus LC

Crocidura fulvastra LC

Crocidura fuscomurina LC

Crocidura russula LC

Crocidura viaria LC

Crocidura voi LC

Ctenodactylus gundi LC

Erythrocebus patas LC

Genetta abyssinica LC

Genetta maculata LC

Genetta pardina LC

Genetta thierryi LC

Helogale parvula LC

Herpestes ichneumon LC

Herpestes sanguineus LC

Hipposideros jonesi NT

Hipposideros vittatus NT

Ichneumia albicauda LC

Ictonyx striatus LC

Mellivora capensis LC

Miniopterus schreibersii NT

Mungos mungo LC

Mustela nivalis LC

316 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Myotis punicus NT

Nyctalus lasiopterus NT

Otomops martiensseni NT

Pachyuromys duprasi LC

Papio hamadryas LC

Papio papio NT

Paraechinus aethiopicus LC

Proteles cristata LC

Rhinolophus euryale NT

Spalax ehrenbergi DD

Sylvicapra grimmia LC

Vulpes cana LC

Vulpes vulpes LC

Xerus erythropus LC

Reptilia Afronatrix anoscopus LC

Afrotyphlops blanfordii DD

Chalcides mionecton LC

Chamaeleo senegalensis LC

Cyrtopodion scabrum LC

Dasypeltis scabra LC

Eumeces algeriensis LC

Gonionotophis grantii LC

Hemidactylus turcicus LC

Hemirhagerrhis hildebrandtii LC

Hemorrhois hippocrepis LC

Lycophidion ornatum LC

Lycophidion semicinctum LC

Macroprotodon brevis NT

Macroprotodon cucullatus LC

Natriciteres olivacea LC

Natrix maura LC

Natrix tessellata LC

Ophisops occidentalis LC

Quedenfeldtia moerens LC FCUP 317 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Saurodactylus brosseti LC

Telescopus variegatus LC H Small-sized amphibians, birds, mammals and reptiles, not endemic, without unique desert adaptations, not threatened, and the majority is carnivorous or omnivorous

Amphibia Conraua beccarii LC

Duttaphrynus dodsoni LC

Hemisus guineensis LC

Hemisus marmoratus LC

Hildebrandtia ornata LC

Hyperolius nitidulus LC

Hyperolius viridiflavus LC

Kassina cassinoides LC

Kassina somalica LC

Leptopelis bufonides LC

Leptopelis viridis LC

Phrynobatrachus LC perpalmatus

Phrynomantis microps LC

Silurana tropicalis LC

Aves Accipiter badius LC

Acrocephalus gracilirostris LC

Acrocephalus rufescens LC

Acrocephalus stentoreus LC

Actophilornis africanus LC

Alaemon alaudipes LC

Alcedo cristata LC

Alectoris barbara LC

Alectoris chukar LC

Alopochen aegyptiaca LC

Amadina fasciata LC

Amandava subflava LC

Amaurornis flavirostra LC

Ammomanes cinctura LC

Ammomanes deserti LC

318 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Ammoperdix heyi LC

Anaplectes rubriceps LC

Anas capensis LC

Anas erythrorhyncha LC

Anas hottentota LC

Anas sparsa LC

Anas undulata LC

Anomalospiza imberbis LC

Anthoscopus parvulus LC

Anthreptes metallicus LC

Anthreptes platurus LC

Anthus leucophrys LC

Anthus similis LC

Apus affinis LC

Apus caffer LC

Apus horus LC

Apus niansae LC

Apus pallidus LC

Batis orientalis LC

Batis senegalensis LC

Bradornis pallidus LC

Bradypterus baboecala LC

Bradypterus cinnamomeus LC

Bubalornis albirostris LC

Bucanetes githagineus LC

Buphagus africanus LC

Buphagus erythrorhynchus LC

Butastur rufipennis LC

Calandrella brachydactyla LC

Calandrella cinerea LC

Calandrella rufescens LC

Camaroptera brachyura LC

Campephaga phoenicea LC

Campethera abingoni LC FCUP 319 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Campethera cailliautii LC

Campethera maculosa LC

Campethera nivosa LC

Campethera nubica LC

Campethera punctuligera LC

Caprimulgus nubicus LC

Carduelis carduelis LC

Carpodacus synoicus LC

Centropus grillii LC

Centropus monachus LC

Centropus senegalensis LC

Centropus superciliosus LC

Cercomela familiaris LC

Cercomela melanura LC

Cercomela scotocerca LC

Cercomela sordida LC

Cercotrichas podobe LC

Certhia brachydactyla LC

Ceyx pictus LC

Charadrius marginatus LC

Charadrius pecuarius LC

Charadrius tricollaris LC

Chersophilus duponti NT

Chlorocichla flavicollis LC

Chrysococcyx cupreus LC

Chrysococcyx klaas LC

Cichladusa guttata LC

Cinnyricinclus leucogaster LC

Cisticola aberrans LC

Cisticola aridulus LC

Cisticola bodessa LC

Cisticola brachypterus LC

Cisticola cantans LC

Cisticola erythrops LC

320 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Cisticola eximius LC

Cisticola galactotes LC

Cisticola guinea LC

Cisticola juncidis LC

Cisticola natalensis LC

Cisticola robustus LC

Cisticola ruficeps LC

Cisticola rufus LC

Cisticola troglodytes LC

Clamator jacobinus LC

Colius striatus LC

Columba livia LC

Coracias abyssinicus LC

Coracias caudatus LC

Coracias cyanogaster LC

Coracias naevius LC

Coracina pectoralis LC

Corvinella corvina LC

Corythaixoides leucogaster LC

Cossypha albicapilla LC

Cossypha heuglini LC

Cossypha niveicapilla LC

Cossypha semirufa LC

Coturnix delegorguei LC

Creatophora cinerea LC

Crecopsis egregia LC

Crinifer piscator LC

Crinifer zonurus LC

Cursorius somalensis LC

Cursorius temminckii LC

Cypsiurus parvus LC

Dendrocopos major LC

Dendrocopos obsoletus LC

Dendropicos fuscescens LC FCUP 321 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Dicrurus adsimilis LC

Dryoscopus gambensis LC

Elanus caeruleus LC

Emberiza affinis LC

Emberiza cia LC

Emberiza flaviventris LC

Emberiza striolata LC

Emberiza tahapisi LC

Empidornis semipartitus LC

Eremalauda dunni LC

Eremomela icteropygialis LC

Eremomela pusilla LC

Eremophila bilopha LC

Eremopterix leucotis LC

Eremopterix nigriceps LC

Erythropygia galactotes LC

Estrilda astrild LC

Estrilda caerulescens LC

Estrilda melpoda LC

Estrilda paludicola LC

Estrilda troglodytes LC

Euplectes afer LC

Euplectes albonotatus LC

Euplectes capensis LC

Euplectes franciscanus LC

Euplectes hordeaceus LC

Eupodotis hartlaubii LC

Eupodotis melanogaster LC

Eupodotis savilei LC

Eupodotis senegalensis LC

Eurystomus glaucurus LC

Falco alopex LC

Falco ardosiaceus LC

Falco chicquera LC

322 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Falco cuvierii LC

Falco pelegrinoides LC

Falco rupicoloides LC

Fringilla coelebs LC

Galerida cristata LC

Galerida modesta LC

Galerida theklae LC

Gallinula angulata LC

Gallinula chloropus LC

Garrulus glandarius LC

Glareola cinerea LC

Glareola nuchalis LC

Glareola pratincola LC

Gypohierax angolensis LC

Halcyon chelicuti LC

Halcyon leucocephala LC

Heliolais erythropterus LC

Himantopus himantopus LC

Hippolais pallida LC

Hirundo abyssinica LC

Hirundo aethiopica LC

Hirundo fuligula LC

Hirundo lucida LC

Hirundo obsoleta LC

Hirundo preussi LC

Hirundo rustica LC

Hirundo semirufa LC

Hirundo senegalensis LC

Hirundo smithii LC

Hyliota flavigaster LC

Indicator indicator LC

Indicator minor LC

Ixobrychus minutus LC

Ixobrychus sturmii LC FCUP 323 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Jynx ruficollis LC

Kaupifalco monogrammicus LC

Lagonosticta larvata LC

Lagonosticta rubricata LC

Lagonosticta rufopicta LC

Lagonosticta senegala LC

Lagonosticta virata LC

Lamprotornis caudatus LC

Lamprotornis chalybaeus LC

Lamprotornis chloropterus LC

Lamprotornis purpureus LC

Lamprotornis purpuroptera LC

Laniarius aethiopicus LC

Laniarius barbarus LC

Laniarius erythrogaster LC

Lanius collaris LC

Lanius excubitoroides LC

Lonchura cantans LC

Lonchura cucullata LC

Lybius bidentatus LC

Lybius dubius LC

Lybius guifsobalito LC

Lybius leucocephalus LC

Lybius rolleti LC

Lybius undatus LC

Malaconotus blanchoti LC

Megaceryle maxima LC

Melaenornis edolioides LC

Melanocorypha calandra LC

Melierax gabar LC

Merops albicollis LC

Merops bulocki LC

Merops hirundineus LC

Merops nubicus LC

324 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Merops orientalis LC

Merops persicus LC

Merops pusillus LC

Merops variegatus LC

Mesopicos goertae LC

Microparra capensis LC

Mirafra africana LC

Mirafra albicauda LC

Mirafra cantillans LC

Mirafra rufa LC

Mirafra rufocinnamomea LC

Motacilla aguimp LC

Motacilla clara LC

Motacilla flava LC

Muscicapa adusta LC

Muscicapa aquatica LC

Muscicapa gambagae LC

Muscicapa striata LC

Myrmecocichla aethiops LC

Myrmecocichla albifrons LC

Myrmecocichla melaena LC

Nectarinia pulchella LC

Nectarinia senegalensis LC

Nettapus auritus LC

Nilaus afer LC

Oena capensis LC

Oenanthe bottae LC

Oenanthe leucopyga LC

Oenanthe leucura LC

Oenanthe lugens LC

Oenanthe moesta LC

Oenanthe monacha LC

Onychognathus morio LC

Oriolus oriolus LC FCUP 325 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Ortygospiza atricollis LC

Ortyxelos meiffrenii LC

Parophasma galinieri LC

Parus ater LC

Parus major LC

Passer eminibey LC

Passer griseus LC

Passer hispaniolensis LC

Passer simplex LC

Passer swainsonii LC

Petronia dentata LC

Petronia petronia LC

Petronia pyrgita LC

Phoeniculus purpureus LC

Phoeniculus somaliensis LC

Phyllolais pulchella LC

Phylloscopus umbrovirens LC

Pica pica LC

Picus vaillantii LC

Pinarocorys erythropygia LC

Platysteira cyanea LC

Plocepasser superciliosus LC

Ploceus badius LC

Ploceus baglafecht LC

Ploceus cucullatus LC

Ploceus galbula LC

Ploceus heuglini LC

Ploceus luteolus LC

Ploceus melanocephalus LC

Ploceus rubiginosus LC

Ploceus taeniopterus LC

Ploceus vitellinus LC

Pluvianus aegyptius LC

Podiceps nigricollis LC

326 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Pogoniulus chrysoconus LC

Pogoniulus pusillus LC

Poicephalus meyeri LC

Poicephalus senegalus LC

Porphyrio alleni LC

Porphyrio porphyrio LC

Porzana pusilla LC

Prinia gracilis LC

Prinia subflava LC

Prionops plumatus LC

Psalidoprocne pristoptera LC

Pseudhirundo griseopyga LC

Pseudoalcippe abyssinica LC

Psittacula krameri LC

Psophocichla litsitsirupa LC

Ptilopachus petrosus LC

Ptilostomus afer LC

Pycnonotus barbatus LC

Pycnonotus xanthopygos LC

Pytilia melba LC

Pytilia phoenicoptera LC

Quelea erythrops LC

Quelea quelea LC

Rallus caerulescens LC

Ramphocoris clotbey LC

Rhinopomastus aterrimus LC

Rhinoptilus chalcopterus LC

Riparia paludicola LC

Riparia riparia LC

Rougetius rougetii NT

Schoenicola brevirostris LC

Scotocerca inquieta LC

Serinus citrinelloides LC

Serinus flavivertex LC FCUP 327 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Serinus gularis LC

Serinus leucopygius LC

Serinus mozambicus LC

Serinus serinus LC

Serinus striolatus LC

Serinus tristriatus LC

Spiloptila rufifrons LC

Stigmatopelia senegalensis LC

Streptopelia capicola LC

Streptopelia decaocto LC

Streptopelia decipiens LC

Streptopelia roseogrisea LC

Streptopelia semitorquata LC

Streptopelia vinacea LC

Sylvia cantillans LC

Sylvia conspicillata LC

Sylvia nana LC

Sylvietta brachyura LC

Tachybaptus ruficollis LC

Tachymarptis aequatorialis LC

Tadorna ferruginea LC

Tchagra minutus LC

Tchagra senegalus LC

Telacanthura ussheri LC

Telophorus sulfureopectus LC

Terpsiphone viridis LC

Tockus erythrorhynchus LC

Tockus fasciatus LC

Tockus flavirostris LC

Tockus hemprichii LC

Tockus nasutus LC

Treron waalia LC

Turdoides leucocephala LC

Turdoides leucopygia LC

328 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Turdoides plebejus LC

Turdoides rubiginosa LC

Turdoides squamiceps LC

Turdus merula LC

Turdus pelios LC

Turnix sylvaticus LC

Turtur abyssinicus LC

Turtur afer LC

Uraeginthus bengalus LC

Urocolius macrourus LC

Vanellus albiceps LC

Vanellus crassirostris LC

Vanellus melanocephalus LC

Vanellus melanopterus LC

Vanellus senegallus LC

Vanellus spinosus LC

Vidua chalybeata LC

Vidua interjecta LC

Vidua larvaticola LC

Vidua macroura LC

Vidua nigeriae LC

Vidua paradisaea LC

Vidua raricola LC

Vidua togoensis LC

Zosterops senegalensis LC

Mammalia Acomys cineraceus LC

Acomys dimidiatus LC

Acomys johannis LC

Acomys russatus LC

Asellia patrizii LC

Atelerix albiventris LC

Atelerix algirus LC

Cardioderma cor LC

Coleura afra LC FCUP 329 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Crocidura aleksandrisi LC

Crocidura cinderella LC

Crocidura foxi LC

Crocidura nanilla LC

Crocidura olivieri LC

Crocidura smithii LC

Crocidura somalica LC

Crocidura whitakeri LC

Crocidura yankariensis LC

Dasymys incomtus LC

Dasymys rufulus LC

Dendromus mystacalis LC

Elephantulus rozeti LC

Eliomys melanurus LC

Eliomys munbyanus LC

Epomophorus gambianus LC

Epomophorus labiatus LC

Eptesicus bottae LC

Eptesicus serotinus LC

Galago senegalensis LC

Gerbilliscus gambiana LC

Gerbilliscus kempi LC

Gerbilliscus robustus LC

Gerbilliscus validus LC

Gerbillus andersoni LC

Gerbillus campestris LC

Gerbillus dasyurus LC

Gerbillus henleyi LC

Gerbillus nanus LC

Gerbillus occiduus LC

Gerbillus simoni LC

Glauconycteris variegata LC

Graphiurus kelleni LC

Graphiurus microtis LC

330 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Heliosciurus gambianus LC

Hemiechinus auritus LC

Hipposideros abae LC

Hipposideros caffer LC

Hipposideros megalotis LC

Hipposideros ruber LC

Lavia frons LC

Lissonycteris angolensis LC

Lophiomys imhausi LC

Lophuromys flavopunctatus LC

Mastomys erythroleucus LC

Mastomys huberti LC

Mastomys natalensis LC

Meriones crassus LC

Meriones grandis LC

Meriones libycus LC

Meriones shawi LC

Micropteropus pusillus LC

Mus haussa LC

Mus mahomet LC

Mus musculus LC

Mus spretus LC

Mus tenellus LC

Mustela subpalmata LC

Myomyscus brockmani LC

Myotis emarginatus LC

Nycteris gambiensis LC

Nycteris hispida LC

Nycteris macrotis LC

Nycteris thebaica LC

Nycticeinops schlieffeni LC

Otonycteris hemprichii LC

Pipistrellus capensis LC

Pipistrellus guineensis LC FCUP 331 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Pipistrellus hesperidus LC

Pipistrellus kuhlii LC

Pipistrellus nanus LC

Pipistrellus rendalli LC

Pipistrellus rueppellii LC

Pipistrellus rusticus LC

Pipistrellus savii LC

Pipistrellus somalicus LC

Plecotus kolombatovici LC

Praomys daltoni LC

Rattus rattus LC

Rhinolophus blasii LC

Rhinolophus clivosus LC

Rhinolophus fumigatus LC

Rhinolophus hipposideros LC

Rhinolophus landeri LC

Rousettus aegyptiacus LC

Scotoecus hirundo LC

Scotophilus dinganii LC

Scotophilus leucogaster LC

Scotophilus nigrita LC

Scotophilus viridis LC

Stenocephalemys albipes LC

Stenocephalemys LC griseicauda

Suncus etruscus LC

Tadarida aegyptiaca LC

Tadarida ansorgei LC

Tadarida bivittata LC

Tadarida condylura LC

Tadarida demonstrator LC

Tadarida major LC

Tadarida midas LC

Tadarida nigeriae LC

332 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Tadarida pumila LC

Tadarida teniotis LC

Taphozous mauritianus LC

Taphozous nudiventris LC

Taphozous perforatus LC

Taterillus congicus LC

Taterillus emini LC

Taterillus gracilis LC

Uranomys ruddi LC

Xerus rutilus LC

Reptilia Acanthocercus annectens LC

Acanthodactylus bedriagai NT

Acanthodactylus maculatus LC

Chalcides chalcides LC

Chalcides polylepis LC

Cynisca feae LC

Duberria lutrix LC

Eirenis coronella LC

Hemidactylus yerburyi LC

Ophisaurus koellikeri LC

Psammodromus algirus LC

Scelarcis perspicillata LC

Tarentola mauritanica LC

Timon pater LC

Timon tangitanus LC

Trachylepis vittata LC

Trogonophis wiegmanni LC I Small-sized diurnal birds and mammals, not endemic, without unique desert adaptations, not threatened, and omnivorous

Aves Cinclus cinclus LC

Dinemellia dinemelli LC

Estrilda rhodopyga LC

Euplectes ardens LC

Euplectes axillaris LC FCUP 333 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Euplectes macroura LC

Halcyon malimbica LC

Lagonosticta rara LC

Lagonosticta rhodopareia LC

Lamprotornis chalcurus LC

Lanius gubernator LC

Monticola rufocinereus LC

Nectarinia cuprea LC

Nectarinia erythrocerca LC

Nectarinia habessinica LC

Nectarinia mariquensis LC

Nectarinia olivacea LC

Nectarinia osea LC

Nectarinia tacazze LC

Nectarinia venusta LC

Onychognathus albirostris LC

Onychognathus blythii LC

Onychognathus tenuirostris LC

Onychognathus tristramii LC

Oriolus monacha LC

Parus leucomelas LC

Parus leuconotus LC

Phyllastrephus strepitans LC

Ploceus superciliosus LC

Prodotiscus regulus LC

Pseudonigrita arnaudi LC

Rhodophoneus cruentus LC

Serinus xanthopygius LC

Sturnus unicolor LC

Sylvia leucomelaena LC

Turdoides reinwardii LC

Turdus olivaceus LC

Zosterops abyssinicus LC

Zosterops poliogastrus LC

334 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Mammalia Crocidura suaveolens LC

Reference International Union for Conservation of Nature (IUCN), 2017. The IUCN Red List of Threatened Species, 2017-3. International Union for Conservation of Nature and Natural Resources. Available at http://www.iucnredlist.org/, Accessed date: 15 November 2017.

FCUP 335 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Appendix B

SUPPLEMENTARY INFORMATION FOR ARTICLE III

336 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Using multivariate statistics to assess ecotourism potential of waterbodies: A case- study in Mauritania.

Frederico Santarém1,2,3, João Carlos Campos Paulo Pereira1,2, Dieng Hamidou5, Jarkko Saarinen4,5, José Carlos Brito1,2

1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 – Faculté des Sciences et Techniques, Université des Sciences, de Technologie et de Médecine de Nouakchott, Nouakchott, Mauritania. 4 – Geography Research Unit, University of Oulu, Finland 5 – School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

List of Supplementary Figures and Tables:

Fig. B1. Results of the ten Bayesian Information Criterion (BIC) models analysed to identify the most likely number of clusters. Fig. B2. Examples of water-bodies in each of the three main groups identified by the Principal Component Analysis. Table B1. Main characteristics of the three groups identified by the PCA and their suitability for different types of ecotourists

FCUP 337 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. B1. Results of the ten Bayesian Information Criterion (BIC) models analysed to identify the most likely number of clusters. Relationships between BIC and number of components are depicted for multiple multivariate models. The model with highest BIC is the spherical, equal volume with 7 components, “EEV”. The BIC increases with the number of clusters and most of the increase occurs with less than seven clusters, which were chosen as the optimum number of clusters defining the groups for ecotourism development. Higher number of clusters brings additional complexity to the analysis without a substantial increment of BIC value (Fraley, Raftery, Murphy, & Scrucca, 2012; Fraley & Raftery, 2002).

338 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Green group

Taghtâfet Blue group

Matmâta Tartêga Red group

Mechaouba Irijil El Gçerba

Fig. B2. Examples of water-bodies in each of the three main groups identified by the Principal Component Analysis. Green group: water-bodies displaying suitable characteristics for soft ecotourists; Blue group: water-bodies suitable for “hard-desert” ecotourists; and Red group: water-bodies less suitable for ecotourism. Note the great surface area in Green water-bodies, the high topography heterogeneity in Blue water-bodies, and the small size of Red water-bodies. See Table B1 for group and water-body characteristics. FCUP 339 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table B1. Main characteristics of the three groups identified by the PCA and their suitability for different types of ecotourists See Figures 2 and 3 for grouping details and geographic distribution of each group, respectively. Groups Water-body Characteristics Suitability Green Boû blei'îne Positive Soft ecotourist Jreif • Numerous flagship Mreimidet El Bidâne species Bougâri • Short distances between Djouk water-bodies Kankossa • High habitat heterogeneity Taghtâfet • Large surface areas Tâmchekket • High water availability • Wide dune cover in surroundings • Wide agriculture cover in surroundings Negative • High temperatures • High human influence index Blue Matmâta Positive Hard ecotourist Tartêga • High topographic Tartêga, upstream of heterogeneity Tkhsutin • High habitat heterogeneity Negative • High temperatures

340 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Red Aouînet Nanâga Positive Less suitable for Ayoun el Khechba • High vegetation diversity ecotourism Laout • High topography Mechaouba heterogeneity Metraoucha Negative Timbâd Ed Dine • Long distances to roads Bednam • Long distances to cities Passes de Diégoum, 10km NW of with supporting Passes de Diégoum, 2km S of infrastructures El Vouk • Numerous IUCN threats El'Atchan El-Khom Sânîyé Irijil El Gçerba Tidâtene Aouinet (palms) Dâber Oumm el Mhâr Oumm Icheglâne

FCUP 341 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Appendix C

SUPPLEMENTARY INFORMATION FOR ARTICLE IV

342 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Mapping and analysing cultural ecosystem services in conflict areas

Frederico Santarém1,2,3, Jarkko Saarinen3,4, José Carlos Brito1,2 1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 - Geography Research Unit, University of Oulu, Finland 4 - School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

List of Supplementary Figures, Tables:

Fig. C1. Sahara-Sahel limits in Africa (small inset) following Dinerstein et al. (2017), range countries within the area, and mountains (yellow balloons), waterbodies (blue balloons), and corridors (red balloons) referred in the text. Adapted from Santarém et al., 2019. BF – Burkina Faso; ER – Eritrea; TU – Tunisia; SEN – Senegal.

Table C1. Types and definitions of the variables chosen to map cultural ecosystem services in the study area.

Table C2. List of flagship species suitable for conservation and ecotourism flagship marketing campaigns in the Sahara-Sahel. Indication of the class, taxa, common name, endemic status (END; Sahara or Sahel), and IUCN conservation status (CS). FCUP 343 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. C1. Sahara-Sahel limits in Africa (small inset) following Dinerstein et al. (2017), range countries within the area, and mountains (yellow balloons), waterbodies (blue balloons), and corridors (red balloons) referred in the text. Adapted from Santarém et al., 2019. BF – Burkina Faso; ER – Eritrea; TU – Tunisia; SEN – Senegal.

References Dinerstein, E., Olson, D., Joshi, A., Vynne, C., Burgess, N.D., Wikramanayake, E., Hahn, N., Palminteri, S., Hedao, P., Noss, R., Hansen, M., Locke, H., Ellis, E.C., Jones, B., Barber, C.V., Hayes, R., Kormos, C., Martin, V., Crist, E., Sechrest, W., Price, L., Baillie, J.E.M., Weeden, D., Suckling, K., Davis, C., Sizer, N., Moore, R., Thau, D., Birch, T., Potapov, P., Turubanova, S., Tyukavina, A., de Souza, N., Pintea, L., Brito, J.C., Llewellyn, O.A., Miller, A.G., Patzelt, A., Ghazanfar, S.A., Timberlake, J., Klöser, H., Shennan-Farpon, Y., Kindt, R., Lillesø, J.-P.B., van Breugel, P., Graudal, L., Voge, M., Al-Shammari, K.F. and Saleem, M. (2017). An ecoregion-based approach to protecting half the terrestrial realm. BioScience, 67: 534-545. doi: 10.1093/biosci/bix014. Santarém, F., Pereira, P., Saarinen, J., Brito, J.C. (2019). New method to identify and map flagship fleets for promoting conservation and ecotourism. Biological Conservation, 229: 113-124. doi: https://doi.org/10.1016/j.biocon.2018.10.017.

344 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table C1. Types and definitions of the variables chosen to map cultural ecosystem services in the study area. Variable Type Definition Major landscape Dune Hill of loose sand built by aeolian processes (wind) or the features flow of water

Plain Vast, open, flat and featureless areas of sandy or gravel substrate

Continental Cliff Significant vertical, or near vertical, rock exposure

Coastal cliff Significant vertical, or near vertical, rock exposure facing the sea

Mountain Significant concentration of canyons, cliffs and escarpments

Plateaux Area of a highland, usually consisting of relatively flat terrain that is raised significantly above the surrounding area, often with one or more sides with steep slopes

Volcanic Crater, cliff, escarpment, inselberg, or plateaux of volcanic origin

Impact Crater Approximately circular depression formed by the hypervelocity impact of a smaller body (confirmed and unconfirmed)

Canyon Deep cleft between escarpments or cliffs resulting from weathering and the erosive activity of a river over geologic timescales

Inselberg Isolated rock hill, knob, ridge, or small mountain that rises abruptly from a gently sloping or virtually level surrounding plain

Valley Significant low area often with a river running through it

Escarpment Steep slope or long rock that occurs from erosion or faulting and separates two relatively level areas of differing elevations Wetlands Permanent Permanent wetland

Seasonal Seasonal wetland

Sebkha Natural salt pan (coastal or continental) where flat expanses of ground are covered with salt and other minerals

River River

Ethnic groups Arab

Beja

Berber

Daza and Ouaddaiens FCUP 345 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Dinka

Dogon

Fulani

Hadendowa

Hausa

Moor

Nubian

Sara and Rounga

Songhai

Soninke

Tuareg

Tubu and Teida

Tukulor

Wolof

Zerma Caravan routes Kanem and Bornu Caravan routes operated by the Kanem and Bornu empires (10th - 19th century)

Ghana, Mali and Caravan routes operated by the , Mali and Songhai Songhai empires (4th - 16th century)

Almoravid and Almohad Caravan routes operated by the Almoravid and Almohad empires (12th - 13th century)

Ancient Egypt Trading route following the Nile river

346 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table C2. List of flagship species suitable for conservation and ecotourism flagship marketing campaigns in the Sahara-Sahel. Indication of the class, taxa, common name, endemic status (END; Sahara or Sahel), and IUCN conservation status (CS). Species data follow IUCN (2019). Adapted from Santarém et al. (2019). Class Taxa Common name END CS Mammalia Addax nasomaculatus Addax Sahara CR Ammotragus lervia Barbary sheep Sahara VU Crocidura tarfayensis Tarfaya Shrew Sahara DD Ctenodactylus vali Val's gundi Sahara DD Equus africanus African wild ass - CR Eudorcas rufifrons Red-fronted gazelle Sahel VU Gazella cuvieri Cuvier's gazelle Sahara VU Gazella dorcas Dorcas gazelle Sahara VU Gazella leptoceros Slender-horned gazelle Sahara EN Giraffa camelopardalis Giraffe - VU Hippopotamus amphibius Common hippopotamus - VU Ictonyx libyca Libyan Striped Weasel Sahara LC Jaculus jaculus Lesser Egyptian jerboa - LC Jaculus orientalis Greater Egyptian jerboa - LC Kobus megaceros Nile lechwe - EN Loxodonta africana African elephant - VU Nanger dama Dama gazelle Sahara CR Nanger soemmerringii Grant's gazelle - VU Oryx beisa Fringe-eared oryx - EN Oryx dammah Scimitar-horned oryx Sahara EW Panthera leo African lion - VU Syncerus caffer African buffalo - NT Trichechus senegalensis African manatee - VU Vulpes zerda Fennec fox Sahara LC Aves Passer cordofanicus Kordofan Sparrow Sahel LC Passer luteus Sudan Golden Sparrow Sahel LC Pterocles alchata Pin-tailed sandgrouse - LC Pterocles coronatus Crowned sandgrouse - LC Pterocles exustus Chestnut-bellied sandgrouse - LC Pterocles gutturalis Yellow-throated sandgrouse - LC Pterocles lichtensteinii Lichtenstein's sandgrouse - LC Pterocles orientalis Black-bellied sandgrouse - LC Pterocles quadricinctus Four-banded sandgrouse - LC Pterocles senegallus Spotted sandgrouse - LC FCUP 347 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Reptilia Acanthodactylus aegyptius Egyptian fringe-fingered lizard Sahara NE Acanthodactylus aureus Golden fringe-fingered lizard Sahara LC Acanthodactylus boskianus Bosc’s fringe-toed lizard - NE Acanthodactylus dumerili Duméril's fringe-fingered lizard Sahara LC Acanthodactylus longipes Long fringe-fingered lizard Sahara NE Acanthodactylus opheodurus Arnold's fringe-fingered lizard - LC Acanthodactylus scutellatus Nidua fringe-fingered lizard - NE Acanthodactylus spinicauda Doumergue's fringe-fingered lizard Sahara CR Acanthodactylus taghitensis Taghit's fringe-toed lizard Sahara DD Agama boueti Bouet's agama Sahel LC Agama boulengeri Boulenger's agama Sahara LC Agama spinosa Spiny Agama - LC Agama tassiliensis Tassili agama Sahara LC Cerastes cerastes Desert horned viper - LC Cerastes vipera Sahara sand viper - LC Chalcides boulengeri Boulenger's feylinia Sahara NE Chalcides delislei De l'Isle's wedge-snouted skink Sahel LC Chalcides humilis Ragazzi's bronze skink Sahara NE Chalcides sepsoides Wedge-snouted skink- Sahara LC Chalcides sphenopsiformis Duméril's wedge-snouted skink Sahara LC Crocodylus suchus West-African crocodile - NE Dasypeltis sahelensis Sahel egg eater Sahel NE Echis coloratus Palestine saw-scaled viper - LC Echis pyramidum Egyptian saw-scaled viper - LC Eryx jaculus Javelin Sand Boa - NE Naja nubiae Nubian spitting cobra Sahel NE Leptotyphlops algeriensis Beaked thread-snake Sahara LC Leptotyphlops boueti Bouet’s worm snake Sahel LC Leptotyphlops cairi Two-coloured blind snake Sahara NE Leptotyphlops macrorhynchus Beaked blind snake - NE Leptotyphlops nursii Nurse's blind snake - NE Mauremys leprosa Mediterranean Turtle - NE Mesalina rubropunctata Red-spotted lizard Sahara NE Pseudocerastes fieldi Field's horned viper - LC Pseudocerastes persicus Perisan horned viper - LC Philochortus lhotei Lhote's shield-backed ground Sahel NE lizard

348 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Pristurus adrarensis Adrar semaphore gecko Sahara DD Pseudotrapelus sinaitus Sinai agama - NE Ptyodactylus guttatus Sinai fan-fingered gecko - NE Ptyodactylus hasselquistii Yellow fan-fingered gecko - NE Ptyodactylus oudrii Oudri's fan-footed gecko - LC Ptyodactylus ragazzi Ragazzi’s fan-footed gecko - NE Ptyodactylus siphonorhina Sinai fan-fingered gecko Sahara NE Python sebae Royal python - NE Scincopus fasciatus Peters' banded skink Sahel DD Scincus albifasciatus Senegal Sandfish Sahara LC Scincus scincus Sandfish skink - NE Stellagama stellio Starred Agama - LC Stenodactylus petri Egyptian sand gecko Sahara NE Stenodactylus stenodactylus Elegant Gecko Sahara NE Testudo kleinmanni Egyptian tortoise Sahara CR Trapelus boehmei Desert agama Sahara LC Trapelus mutabilis Desert agama Sahara NE Trapelus pallidus Pallid agama - NE Trapelus schmitzi Schmitz’ agama Sahara DD Trapelus tournevillei Sahara agama Sahara LC Tropiocolotes algericus Algerian sand gecko Sahara LC Tropiocolotes bisharicus Bishari pigmy gecko Sahara NE Tropiocolotes steudeneri Algerian sand gecko Sahara NE Typhlops etheridgei Mauritanian Blind Snake Sahara DD Typhlops vermicularis European blind snake - LC Uromastyx acanthinura Schmidt’s spiny-tailed lizard Sahara NE Uromastyx aegyptia Egyptian spiny–tailed lizard - VU Uromastyx alfredschmidti Schmidt’s mastigure Sahara NT Uromastyx dispar Sudan mastigure Sahara NE Uromastyx geyri Geyr’s spiny-tailed lizard Sahara NT Uromastyx nigriventris Moroccan Spiny-tailed Lizard Sahara NE Uromastyx occidentalis Giant spiny-tailed lizard Sahara NE Uromastyx ocellate Ocellated Spinytail - NE Uromastyx ornate Ornate mastigure - NE Varanus griseus Desert monitor - NE

References FCUP 349 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

International Union for Conservation of Nature (IUCN). (2019). The IUCN Red List of Threatened Species, 2017-3. International Union for Conservation of Nature and Natural Resources. Accessed 02 June 2019 at http://www.iucnredlist.org/. Santarém, F., Pereira, P., Saarinen, J., Brito, J.C. (2019). New method to identify and map flagship fleets for promoting conservation and ecotourism. Biological Conservation, 229: 113-124. doi: https://doi.org/10.1016/j.biocon.2018.10.017.

350 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Appendix D

SUPPLEMENTARY INFORMATION FOR ARTICLE V FCUP 351 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Assessment and prioritization of cultural ecosystem services in the Sahara-Sahelian Frederico Santarém1,2,3, Jarkko Saarinen3,4, José Carlos Brito1,2 1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 - Geography Research Unit, University of Oulu, Finland 4 - School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

List of Supplementary Figures, Tables: and Text:

Fig. D1 – Sahara-Sahel limits in Africa (small inset), its range countries (three-letter codes), the main mountains (yellow balloons), and wetlands (blue balloons) are depicted. ALG: Algeria; BUF: Burkina Faso; CAM: Cameroon; CHA: Chad; EGY: Egypt; ERI: Eritrea; ETH: Ethiopia; LIB: Libya; MAL: Mali; Mauritania: MAU; MOR: Morocco; NIG: Niger; NIA: Nigeria; SEN: Senegal; SSU: South Sudan; SUD: Sudan; and TUN: Tunisia. Figure adapted from Santarém et al. (2020).

Fig. D2 – Relationship between the number of inbound tourists in Sahara-Sahel countries and their rank in the Human Development Index Relationship between the number of inbound tourists (UNWTO, 2019) in Sahara-Sahel countries and their rank in the Human Development Index (HDI) (UNDP, 2019). Tourism data are for 2017 and are represented in millions. No data are available for South Sudan and thus are not represented in the graph.

Fig. D3 – Analytical workflow used to evaluate the performance in managing CES among countries within Sahara-Sahel. The framework combines spatially explicit features and costs (point, polygon, and polyline data at 0.5º resolution) used to prioritize the landscape (A) with predictors of performance in preserving cultural ecosystem services among countries within Sahara-Sahel ecoregions (B). The framework produces spatially and statistically explicit outputs from which it can be identified which countries need to improve touristic, economic, governance, environmental, health, security, and socioeconomic conditions to attract additional international tourists and which ones are missing opportunities given the high potential of their ecosystems to supply cultural services. See Tables 1 and 3 for a detailed explanation of categories used in the prioritization rankings and in the statistical analyses, respectively.

352 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D4 – Sensitivity analyses for different cost weightings (1x, 100x, and 1000x) on 10 different ranking scenarios (a-j). Pixels (0.5º resolution) are colored by ranking priorities of CES availability: blue – low; yellow – middle; and red – high priority for CES conservation. See Table 2 for a detailed description of the 10 scenarios.

D5 – Spatial variability of the high prioritized cells for CES conservation in the Sahara-Sahel. The composite map represents the two first components (PC1 – green, PC2 – blue) of the spatial PCA (89% of the variance explained), which are colored according to the availability of CES (PC1 – green; high priority; PC2 – blue; low priority) at 0.5º resolution. See Fig. 4.1 for the frequency of selection of high priority rankings.

Fig. D6 – Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socioeconomic indicators are constraining national conditions for CES conservation (Fig. 4.3 and Table 4.4). See Fig. D1 for country codes

Fig. D7 – Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation. Conditions measured by the Human Development Index (HDI) (UNDP, 2019). See Fig. D1 for country codes

Fig. D8 – Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation and the Zonation prioritization ranking. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socio-economic indicators are constraining national conditions for CES (Fig. 4.3 and Table 4.4). See Fig. 4.1 for Zonation rankings. See Fig. D1 for country codes.

Fig. D9 – Sensitivity analysis of the relationship between the conditions for visitation and the availability of CES in each country considering the Zonation prioritization ranking. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socio-economic indicators are constraining national conditions for CES (Fig. 4.3 and Table 4.4). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country)) highly ranked for CES conservation. See Fig. D1 for Zonation rankings and Fig. D1 for country codes. FCUP 353 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D10 – Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation and the Zonation prioritization ranking. Conditions were measured by the Human Development Index (UNDP, 2019). Thresholds of medium developed countries (0.55) and high developed countries (0.70) are marked with a dashed line. See Fig. 4.1 for Zonation rankings and Fig. D1 for country codes.

Fig. D11 – Sensitivity analysis of the relationship between the conditions for visitation and the availability of CES in each country considering the Zonation prioritization rankings. Conditions for visitation were retrieved from the Human Development Index (HDI) (UNDP, 2019). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country)) highly ranked for CES conservation. Thresholds of medium developed countries (0.55) and high developed countries (0.70) are marked with a dashed line. See Fig. D1 for country codes.

Table D1 – Descriptive statistics of socio-economic variables (N=17) for the Sahara-Sahel countries (N=18). Minimum (Min), maximum (Max), mean and the standard deviation (SD) are shown. See Table 3 for variables codes.

Table D2 – The squared cosines of the variables on the principal components (PCs). The greater the squared cosine value, the greater the link with the corresponding axis. Values in bold correspond for each variable to the factor for which the squared cosine is the largest. For simplicity, only the six first PCs are shown. See Table 4.3 for variables codes and Table 4.4 for other PCA results.

Text D1 – Detailed explanation on the spatial data acquisition and processing

Text D2 – Explanation of chosen variables for statistical analyses

354 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D1. Sahara-Sahel limits in Africa (small inset), its range countries (three-letter codes), the main mountains (yellow balloons), and wetlands (blue balloons) are depicted. ALG: Algeria; BUF: Burkina Faso; CAM: Cameroon; CHA: Chad; EGY: Egypt; ERI: Eritrea; ETH: Ethiopia; LIB: Libya; MAL: Mali; Mauritania: MAU; MOR: Morocco; NIG: Niger; NIA: Nigeria; SEN: Senegal; SSU: South Sudan; SUD: Sudan; and TUN: Tunisia. Figure adapted from Santarém et al. (2020).

FCUP 355 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D2. Relationship between the number of inbound tourists (UNWTO, 2019) in Sahara-Sahel countries and their rank in the Human Development Index (HDI) (UNDP, 2019). Tourism data are for 2017 and are represented in millions. No data are available for South Sudan and thus are not represented in the graph.

356 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D3. Analytical workflow used to evaluate the performance in managing CES among countries within Sahara-Sahel. The framework combines spatially explicit features and costs (point, polygon, and polyline data at 0.5º resolution) used to prioritize the landscape (A) with predictors of performance in preserving cultural ecosystem services among countries within Sahara-Sahel ecoregions (B). The framework produces spatially and statistically explicit outputs from which it can be identified which countries need to improve touristic, economic, governance, environmental, health, security, and socioeconomic conditions to attract additional international tourists and which ones are missing opportunities given the high potential of their ecosystems to supply cultural services. See Tables 1 and 3 for a detailed explanation of categories used in the prioritization rankings and in the statistical analyses, respectively.

FCUP 357 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D4. Sensitivity analyses for different cost weightings (1x, 100x, and 1000x) on 10 different ranking scenarios (a-j). Pixels (0.5º resolution) are colored by ranking priorities of CES availability: blue – low; yellow – middle; and red – high priority for CES conservation. See Table 2 for a detailed description of the 10 scenarios.

358 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D5. Spatial variability of the high prioritized cells for CES conservation in the Sahara-Sahel. The composite map represents the two first components (PC1 – green, PC2 – blue) of the spatial PCA (89% of the variance explained), which are colored according to the availability of CES (PC1 – green; high priority; PC2 – blue; low priority) at 0.5º resolution. See Fig. 4.1 for the frequency of selection of high priority rankings.

FCUP 359 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D6. Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socioeconomic indicators are constraining national conditions for CES conservation (Fig. 4.3 and Table 4.4). See Fig. D1 for country codes

360 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D7. Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation. Conditions measured by the Human Development Index (HDI) (UNDP, 2019). See Fig. D1 for country codes.

FCUP 361 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D8. Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation and the Zonation prioritization ranking. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socio-economic indicators are constraining national conditions for CES (Fig. 4.3 and Table 4.4). See Fig. 4.1 for Zonation rankings. See Fig. D1 for country codes.

362 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D9. Sensitivity analysis of the relationship between the conditions for visitation and the availability of CES in each country considering the Zonation prioritization ranking. Conditions were retrieved from the PC1 loadings of the PCA performed for assessing which socio-economic indicators are constraining national conditions for CES (Fig. 4.3 and Table 4.4). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country)) highly ranked for CES conservation. See Fig. D1 for Zonation rankings and Fig. D1 for country codes.

FCUP 363 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D 10. Relationship between the conditions for visitation and the availability of CES in each country, considering the raw number of planning units – PUs (N) – highly ranked for CES conservation and the Zonation prioritization ranking. Conditions were measured by the Human Development Index (UNDP, 2019). Thresholds of medium developed countries (0.55) and high developed countries (0.70) are marked with a dashed line. See Fig. 4.1 for Zonation rankings and Fig. D1 for country codes.

364 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. D 11. Sensitivity analysis of the relationship between the conditions for visitation and the availability of CES in each country considering the Zonation prioritization rankings. Conditions for visitation were retrieved from the Human Development Index (HDI) (UNDP, 2019). Availability of CES was measured by the number of planning units weighted by the area of the Sahara-Sahel ecoregions (Dinerstein et al., 2017) within each country (PUs (N/country)) highly ranked for CES conservation. Thresholds of medium developed countries (0.55) and high developed countries (0.70) are marked with a dashed line. See Fig. D1 for country codes.

FCUP 365 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table D1. Descriptive statistics of socio-economic variables (N=17) for the Sahara-Sahel countries (N=18). Minimum (Min), maximum (Max), mean and the standard deviation (SD) are shown. See Table 3 for variables codes.

Min. Max. Mean SD TOU 30000 11349000 2112752 3283392 VIS 1 95 20.444 25.978 GNI 920 21340 5802.222 5732.136 MPI 0.1 74.8 35.789 26.955 GEF 0.0 48.560 20.942 16.148 BUS 53 189 147.889 37.818 CON 8 158 67.061 44.996 BIO -3.1 6.5 -0.022 1.965 THR 34 888 171.833 200.044 HIV 0.1 3.6 1.000 1.149 TUB 12 540 122.111 117.770 WAS 0.6 82.1 35.383 28.590 INT 1.31 64.80 27.396 21.371 ELE 10.9 100.0 55.256 29.107 GPI 1.883 3.526 2.574 0.464 GTI 0.0 8.6 4.722 2.563 ROA 9.7 33.6 25.011 5.178

366 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table D2. The squared cosines of the variables on the principal components (PCs). The greater the squared cosine value, the greater the link with the corresponding axis. Values in bold correspond for each variable to the factor for which the squared cosine is the largest. For simplicity, only the six first PCs are shown. See Table 4.3 for variables codes and Table 4.4 for other PCA results.

PC1 PC2 PC3 PC4 PC5 PC6

TOU 0.607 0.000 0.136 0.010 0.050 0.064 VIS 0.391 0.174 0.122 0.061 0.046 0.073 GNI 0.397 0.444 0.010 0.017 0.033 0.060 MPI 0.610 0.149 0.020 0.168 0.006 0.004 GEF 0.634 0.131 0.053 0.043 0.023 0.034 BUS 0.634 0.086 0.115 0.115 0.006 0.011 CON 0.149 0.359 0.067 0.193 0.000 0.115 BIO 0.426 0.176 0.131 0.087 0.113 0.002 THR 0.004 0.043 0.390 0.071 0.454 0.011 HIV 0.506 0.005 0.384 0.014 0.012 0.000 TUB 0.368 0.011 0.390 0.042 0.107 0.012 WAS 0.605 0.068 0.049 0.050 0.002 0.051 INT 0.811 0.004 0.086 0.002 0.002 0.001 ELE 0.762 0.078 0.057 0.051 0.003 0.002 GPI 0.419 0.338 0.002 0.003 0.121 0.036 GTI 0.143 0.412 0.087 0.250 0.000 0.018 ROA 0.554 0.022 0.005 0.011 0.007 0.300

FCUP 367 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Text D1. Detailed explanation on the spatial data acquisition and processing

A set of 148 spatially explicit and fine-scale data layers representative of 24 variables influencing CES in deserts were collected from multiple sources (see sources in Table 4.1). The layers consisted of positive and cost features to the prioritization ranking. Methodological details on the acquisition and process of these data are provided here. Species data (2 variables): we retrieved species range data from the two main groups of large- and small-bodied flagship species in Sahara-Sahel (including mammals, birds, and reptiles), as they are the most suitable candidates for conservation and ecotourism promotion in the region (Santarém et al., 2019). Distribution data for these flagship species (and for many of the other features; see below) were available as polygons representing the extent of occurrence where species (or other features) occur. These distribution polygons may overestimate features’ ranges (Graham and Hijmans, 2006; Rondinini et al., 2006) or, in the case of Sahara-Sahel, underestimate due to scarce sampling efforts (Brito et al., 2014a, 2016). Yet, the collated range maps represent the most advanced and updated information available on the distribution of the features considered in this work (Santarém et al., 2020a). Other biodiversity data (3 variables): we took data on the distribution of forest reserves, protected areas and UNESCO World Heritage Sites (Santarém et al., 2020a), as these areas are key factors for CES (Balmford et al., 2015; Buckley et al., 2019; Naidoo et al., 2019). Landscape data (8 variables): We collated data on major landscape features, gorges and mountain passes, peculiar rock formations, caves, major wetlands, and rock pools (Santarém et al., 2020a), as these desert landscape traits provide many CES (Santarém et al., 2020b) and are key factors for desert biodiversity (Vale et al., 2015). We also pooled data on desert ecosystem intactness (low Human Footprint values) and areas of extreme remoteness (Venter et al., 2018, 2016a, 2016b), as isolated, inaccessible and detached wild places (Di Marco et al., 2019; Watson et al., 2018) have been associated with high levels of human mental health and wellbeing (Bratman et al., 2019; Buckley et al., 2019; Dhami et al., 2014; Kay Smith and Diekmann, 2017; Naidoo et al., 2019). The Human Footprint is the most complete terrestrial dataset of high- resolution cumulative human pressures on the environment ever developed (Venter et al., 2018). Cultural data (7 variables): we gathered cultural-related features, such as Sahara-Sahel ethnographic groups, oases, monuments, caravan villages, fortifications from the colonial period, sites of historical occupation and rock art (Santarém et al., 2020a), as these features add value to desert ecosystems and are vital assets to CES in drylands (Davies et al., 2012; Levin et al., 2019; Santarém et al., 2020b; UNEP, 2006a, 2006b).

368 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Conflict data (2 variables): we collected conflict and landmines data from Santarém et al., (2020a), given that the success of prioritization exercises is constrained by the need to consider complex and problematic land uses during the planning stage (Moilanen et al., 2011; Naidoo et al., 2006). Widespread factors, such as armed conflicts, may complicate priority-setting efforts (Daskin and Pringle, 2018; Hammill et al., 2016), and thus need to enter as costs in systematic planning (Hammill et al., 2016; Moilanen et al., 2014; Naidoo et al., 2006). Exploitation of natural resources data (2 variables): Overexploitation of natural resources is dramatically decreasing the already imperiled desert diversity, compromising the supply of ecosystem services even further (Brito et al., 2014a; Cerretelli et al., 2018; Duncan et al., 2014; Taylor et al., 2017). We collated data of exploitation facilities in Sahara-Sahel from a recent work (Santarém et al., 2020a). Contrary to extreme remote areas, areas with high cumulative human pressures tend to supply less CES (Santarém et al., 2020a; Venter et al., 2016a). We thus pooled data related to high levels of Human Footprint (Venter et al., 2018, 2016a, 2016b).

FCUP 369 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Text D2. Explanation of chosen variables for statistical analyses

A set of 15 socio-economic data representative of country-conditions were collected from multiple sources (see sources in Table 4.3). Methodological details on the acquisition and process of these data are provided here.

Tourism data (2 variables): Tourism levels are often associated with better conditions for visiting (UNWTO, 2019). Particularly, potential tourists tend to prefer countries in which the ease of mobility is superior (Whyte, 2008). We recorded the number of international tourist arrivals (which are not limited to, but include ecotourists) (TOU). International inbound tourists (overnight visitors) represent the number of tourists who travel to a country other than that in which they usually reside, and outside their usual environment, for a period not exceeding 12 months and whose main purpose in visiting is other than an activity remunerated from within the country visited. When data on number of tourists are not available, the number of visitors, which includes tourists, same- day visitors, cruise passengers, and crew members, is shown instead (UNWTO, 2019). Information on visit rates may be confounded by variation in methods (Balmford et al., 2015), but we used the most updated database of visit rates (UNWTO, 2019), which has reduced this limitation. We also accessed the ease of movement between countries by counting the number of VISA-free passports that each country accepts at their borders (VIS). Passport Index tracks the world’s ease of mobility. Rankings are based on the freedom of movement and visa-free travel open to holders of passports (Passport Index, 2019).

Economic data (2 variables): economic levels dictate expansions or reductions in conservation spending (Bradshaw and Di Minin, 2019; Waldron et al., 2017, 2013). We followed previous authors in using national rates for economic growth (e.g., Balmford et al., 2015; Bradshaw et al., 2010; Bradshaw and Di Minin, 2019; Waldron et al., 2017). We accessed the gross national income per capita corrected for purchasing-power parity (GNI) (World Bank, 2019). GNI per capita based on purchasing power parity (PPP) is the gross national income (GNI) converted to international dollars using PPP rates. An international dollar has the same purchasing power over GNI as a U.S. dollar has in the United States. GNI is the sum of value added by all resident producers plus any product taxes (less subsidies) not included in the valuation of output plus net receipts of primary income (compensation of employees and property income) from abroad. Data are in current international dollars based on the 2011 International Comparison Program round

370 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

(World Bank, 2019).We also accounted for the percentage of the national population that is multidimensionally poor, which adjusts poverty to the intensity of deprivations and reveals inequalities across and within countries (MPI). The global multidimensional poverty index denotes the percentage of the population that is multidimensionally poor adjusted by the intensity of the deprivations. It sheds light on the number of people experiencing poverty at regional, national, and subnational levels, and reveal inequalities across countries and among the poor themselves (OPHI and UNDP, 2019). Governance data (2 variables): Good governance positively affects tourism businesses (Hausmann et al., 2017a) and ecosystem conservation efficiency (Bradshaw and Di Minin, 2019; Waldron et al., 2017). There are many indicators to measure governance (Kaufmann et al., 2011); yet, government effectiveness (GEF) is considered the best indicator to measure the performance of countries in achieving conservation management goals (Waldron et al., 2017). GEF reflects perceptions of the quality of public services, the quality of the civil service and the degree of its independence from political pressures, the quality of policy formulation and implementation, and the credibility of the government's commitment to such policies (Kaufmann et al., 2011). We collected data on the GEF in each country, which reflects the degree of independence from political pressures and the quality policy formulation and implementation (Kaufmann et al., 2011). We also collected data on the regulatory environment of each country to conduct business operations (BUS) from the World Bank statistical tables (World Bank, 2019). Ease of doing business is a numerical rank of the world’s countries concerning the regulatory environment in which a business operation is conducted. The index averages the country's percentile rankings on 10 topics covered in the World Bank's Doing Business. The ranking on each topic is the simple average of the percentile rankings on its component indicators (World Bank, 2019). Environmental data (3 variables): Losses in ecosystem services and species conservation up- listing have been associated to inadequate funding (Powers and Jetz, 2019; Waldron et al., 2017, 2013). Additionally, the capacity of ecosystems to regenerate what people demand from them is threatened by the overexploitation of natural resources (Cardinale et al., 2012), especially in deserts (Brito et al., 2014a; Díaz et al., 2019; Duncan et al., 2014; Santarém et al., 2020a). We took data on average annual conservation investment levels (CON) from Waldron et al. (2013), which collated country-level conservation funding flowing from a broad range of trust funds, international donors, domestic governments, and self-funding schemes via user payments. We collected data on the biocapacity of each country (in hectares) to regenerate its ecosystems (BIO), a similar indicator to the Ecological Footprint (Global Footprint Network, 2019; Lin et al., 2018). Biocapacity is the capacity of ecosystems to regenerate what people demand from those FCUP 371 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel surfaces. The biocapacity of a surface represents its ability to renew what people demand. Biocapacity is therefore the ecosystems' capacity to produce biological materials used by people and to absorb waste material generated by humans, under current management schemes and extraction technologies. Biocapacity can change from year to year due to climate, management, and proportion considered useful inputs to the human economy. Biocapacity is expressed in global hectares. We used the IUCN Red List of Threatened Species (IUCN, 2019) to account for the number of threatened vertebrates, mollusks, other invertebrates, plants, fungi and protists in each country (THR). The List gives a critical indicator of the health of the world’s biodiversity and helps catalyzing action for biodiversity conservation and policy change, critical to protecting the natural resources needed for survival. It provides information about range, population size, habitat, and ecology, use and/or trade, threats, and conservation actions that will help inform necessary conservation decisions (IUCN, 2019).

Health data (3 variables): health is a major concern to tourists travelling abroad, especially in underdeveloped countries (Jonas et al., 2010). Tuberculosis and other infections or diseases that can be grabbed from untreated waters are reasons of concern to travelers in tropical countries and may compromise their demand for CES (Jonas et al., 2010). We collected data on the incidence of tuberculosis per 100,000 people (TUB), which comprises the estimated number of new and relapse tuberculosis cases arising each year, expressed as the rate per 100,000 population. All forms of tuberculosis are included, including cases in people living with HIV. Estimates for all years are recalculated as new information becomes available and techniques are refined, so they may differ from those published previously (World Bank, 2019; World Health Organization, 2019). We also collected data related to deaths attributable to unsafe water, sanitation, and hygiene (WAS) focusing on inadequate WAS services per 100,000 population. Death rates are calculated by dividing the number of deaths by the total population. In this estimate, only the impact of diarrheal diseases, intestinal nematode infections, and protein-energy malnutrition are considered (WAS) (World Bank, 2019; World Health Organization, 2019). We also collected data on the total prevalence of the human immunodeficiency virus (HIV) as a percentage of the population ages 15-49 who are infected with the HIV (World Bank, 2019).

Social data (2 variables): Accessibility to electricity and to internet influence visitors’ demand for CES. In particular, internet accessibility has been correlated with tourist’s satisfaction when they are enjoying CES, as they likely to record and share pictures of local biodiversity, landscapes and ecosystems in social media (Hausmann et al., 2017b, 2017a). We collected data on the

372 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

percentage of population with access to electricity (ELE) (electrification data are collected from industry, national surveys, and international sources) and to internet (with Internet users considered those individuals who have used the Internet in the last 3 months from a computer, mobile phone, digital TV, or other electronic equipment) (INT) from the World Bank statistical tables (World Bank, 2019). Security data (3 variables): As insecurity and terrorism events are threatening people and ecosystems in Sahara-Sahel (Brito et al., 2018, 2014b; Durant and Brito, 2019; Santarém et al., 2020a), we collected data on the country-based state of peace from the Global Peace Index (GPI). The index ranks 163 countries and territories according to their level of peacefulness using 23 qualitative and quantitative indicators from highly respected sources and measures the state of peace using three thematic domains: the level of Societal Safety and Security, the extent of Ongoing Domestic and International Conflict, and the degree of Militarization (Institute for Economics & Peace, 2019a). We thus collated patterns of terrorism (GTI) from the Global Terrorism Index. The index ranks 138 countries and provides a comprehensive summary of the key global trends and patterns in terrorism over the last 50 years, covering the period from the beginning of 1970 to the end of 2018, and placing a special emphasis on trends since 2014, which corresponds with the start of the fall of Islamic State of Iraq and the Levant (Institute for Economics & Peace, 2019b). Finally, we collected mortality data related to traffic injuries (ROA), which measures the estimated road traffic fatal injury deaths per 100,000 population (World Health Organization, 2019). FCUP 373 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

References Balmford, A., Green, J.M., Anderson, M., Beresford, J., Huang, C., Naidoo, R., Walpole, M., Manica, A., 2015. Walk on the wild side: Estimating the global magnitude of visits to protected areas. PLoS Biol. 13, e1002074. https://doi.org/10.1371/journal.pbio.1002074 Bradshaw, C.J.A., Di Minin, E., 2019. Socio-economic predictors of environmental performance among African nations. Sci. Rep. 9, 1–13. https://doi.org/10.1038/s41598-019-45762-3 Bradshaw, C.J.A., Giam, X., Sodhi, N.S., 2010. Evaluating the relative environmental impact of countries. PLoS One 5. https://doi.org/10.1371/journal.pone.0010440 Bratman, G.N., Anderson, C.B., Berman, M.G., Cochran, B., de Vries, S., Flanders, J., Folke, C., Frumkin, H., Gross, J.J., Hartig, T., Kahn, P.H., Kuo, M., Lawler, J.J., Levin, P.S., Lindahl, T., Meyer-Lindenberg, A., Mitchell, R., Ouyang, Z., Roe, J., Scarlett, L., Smith, J.R., van den Bosch, M., Wheeler, B.W., White, M.P., Zheng, H., Daily, G.C., 2019. Nature and mental health: An ecosystem service perspective. Sci. Adv. 5. https://doi.org/10.1126/sciadv.aax0903 Brito, J.C., Durant, S.M., Pettorelli, N., Newby, J., Canney, S., Algadafi, W., Rabeil, T., Crochet, P.-A., Pleguezuelos, J.M., Wacher, T., Smet, K. De, Gonçalves, D.V., da Silva, M.J.F., Martínez-Freiría, F., Abáigar, T., Campos, J.C., Comizzoli, P., Fahd, S., Fellous, A., Garba, H.H.M., Hamidou, D., Harouna, A., Hatcha, M.H., Nagy, A., Silva, T.L., Sow, A.S., Vale, C.G., Boratyński, Z., Rebelo, H., Carvalho, S.B., 2018. Armed conflicts and wildlife decline : Challenges and recommendations for effective conservation policy in the Sahara-Sahel. Conserv. Lett. 11, 1–13. https://doi.org/10.1111/conl.12446 Brito, J.C., Godinho, R., Martínez-Freiría, F., Pleguezuelos, J.M., Rebelo, H., Santos, X., Vale, C.G., Velo-Antón, G., Boratyński, Z., Carvalho, S.B., Ferreira, S., Gonçalves, D. V., Silva, T., Tarroso, P., Campos, J.C., Leite, J. V., Nogueira, J., Alvares, F., Sillero, N., Sow, A.S., Fahd, S., Crochet, P.-A., Carranza, S., 2014a. Unravelling biodiversity, evolution and threats to conservation in the Sahara-Sahel. Biol. Rev. 89, 215–231. https://doi.org/10.1111/brv.12049 Brito, J.C., Godinho, R., Martínez-Freiría, F., Pleguezuelos, J.M., Rebelo, H., Santos, X., Vale, C.G., Velo-Antón, G., Boratyński, Z., Carvalho, S.B., Ferreira, S., Gonçalves, D. V., Silva, T.L., Tarroso, P., Campos, J.C., Leite, J. V., Nogueira, J., Alvares, F., Sillero, N., Sow, A.S., Fahd, S., Crochet, P.-A.A., Carranza, S., Álvares, F., Sillero, N., Sow, A.S., Fahd, S., Crochet, P.-A.A., Carranza, S., 2014b. Unravelling biodiversity, evolution and threats to conservation in the sahara-sahel. Biol. Rev. 89, 215–231. https://doi.org/10.1111/brv.12049 Brito, J.C., Tarroso, P., Vale, C.G., Martínez-Freiría, F., Boratyński, Z., Campos, J.C., Ferreira,

374 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

S., Godinho, R., Gonçalves, D. V., Leite, J. V., Lima, V.O., Pereira, P., Santos, X., da Silva, M.J.F.. F., Silva, T.L., Velo-Antón, G., Veríssimo, J., Crochet, P.A., Pleguezuelos, J.M., Carvalho, S.B., 2016. Conservation biogeography of the Sahara-Sahel: Additional protected areas are needed to secure unique biodiversity. Divers. Distrib. 22, 371–384. https://doi.org/10.1111/ddi.12416 Buckley, R., Brough, P., Hague, L., Chauvenet, A., Fleming, C., Roche, E., Sofija, E., Harris, N., 2019. Economic value of protected areas via visitor mental health. Nat. Commun. 10, 1–10. https://doi.org/10.1038/s41467-019-12631-6 Cardinale, B.J., Duffy, J.E., Gonzalez, A., Hooper, D.U., Perrings, C., Venail, P., Narwani, A., MacE, G.M., Tilman, D., Wardle, D.A., Kinzig, A.P., Daily, G.C., Loreau, M., Grace, J.B., Larigauderie, A., Srivastava, D.S., Naeem, S., 2012. Biodiversity loss and its impact on humanity. Nature 486, 59–67. https://doi.org/10.1038/nature11148 Cerretelli, S., Poggio, L., Gimona, A., Yakob, G., Boke, S., Habte, M., Coull, M., Peressotti, A., Black, H., 2018. Spatial assessment of land degradation through key ecosystem services: The role of globally available data. Sci. Total Environ. 628–629, 539–555. https://doi.org/10.1016/j.scitotenv.2018.02.085 Daskin, J.H., Pringle, R.M., 2018. Warfare and wildlife declines in Africa’s protected areas. Nature 553, 328–332. https://doi.org/10.1038/nature25194 Davies, J., Poulsen, L., Schulte-Herbrüggen, B., Mackinnon, K., Crawhall, N., Hendwood, W.D., Dudley, N., Smith, J., Gudka, M., 2012. Conserving Dryland Biodiversity. International Union for the Conservation of Nature, Nairobi. Dhami, I., Deng, J., Burns, R.C., Pierskalla, C., 2014. Identifying and mapping forest-based ecotourism areas in West Virginia - Incorporating visitors’ preferences. Tour. Manag. 42, 165–176. https://doi.org/10.1016/j.tourman.2013.11.007 Di Marco, M., Ferrier, S., Harwood, T.D., Hoskins, A.J., Watson, J.E.M., 2019. Wilderness areas halve the extinction risk of terrestrial biodiversity. Nature 573, 582–585. https://doi.org/https://doi.org/10.1038/s41586-019-1567-7 Díaz, S., Settele, J., Brondízio, E.S., Ngo, H.T., Agard, J., Arneth, A., Balvanera, P., Brauman, K.A., Butchart, S.H.M., Chan, K.M.A., Lucas, A.G., Ichii, K., Liu, J., Subramanian, S.M., Midgley, G.F., Miloslavich, P., Molnár, Z., Obura, D., Pfaff, A., Polasky, S., Purvis, A., Razzaque, J., Reyers, B., Chowdhury, R.R., Shin, Y.J., Visseren-Hamakers, I., Willis, K.J., Zayas, C.N., 2019. Pervasive human-driven decline of life on Earth points to the need for transformative change. Science (80-. ). 366, eaax3100. https://doi.org/10.1126/science.aax3100 FCUP 375 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Dinerstein, E., Olson, D., Joshi, A., Vynne, C., Burgess, N.D., Wikramanayake, E., Hahn, N., Palminteri, S., Hedao, P., Noss, R., Hansen, M., Locke, H., Ellis, E.C., Jones, B., Barber, C.V., Hayes, R., Kormos, C., Martin, V., Crist, E., Sechrest, W.E.S., Price, L., Baillie, J.E.M., Weeden, D.O.N., Suckling, K., Davis, C., Sizer, N., Moore, R., Thau, D., Birch, T., Potapov, P., Turubanova, S., Tyukavina, A., Souza, N.D.E., Pintea, L., Brito, J.C., Timberlake, J., Klöser, H., Shennan-farpón, Y., Kindt, R., 2017. An ecoregion-based approach to protecting half the terrestrial realm. Bioscience 67, 534–545. https://doi.org/10.1093/biosci/bix014 Duncan, C., Kretz, D., Wegmann, M., Rabeil, T., Pettorelli, N., 2014. Oil in the Sahara: Mapping anthropogenic threats to Saharan biodiversity from space. Philos. Trans. R. Soc. B 369, 20130191. https://doi.org/http://dx.doi.org/10.1098/rstb.2013.0191 Durant, S.M., Brito, J.C., 2019. Protect environment from armed conflicts. Nature 571, 478. Global Footprint Network, 2019. Global Footprint Network National Footprint Accounts, 2019 Edition [WWW Document]. URL http://data.footprintnetwork.org (accessed 1.19.20). Graham, C.H., Hijmans, R.J., 2006. A comparison of methods for mapping species ranges and species richness. Glob. Ecol. Biogeogr. 15, 578–587. https://doi.org/10.1111/j.1466- 8238.2006.00257.x Hammill, E., Tulloch, A.I.T., Possingham, H.P., Strange, N., Wilson, K.A., 2016. Factoring attitudes towards armed conflict risk into selection of protected areas for conservation. Nat. Commun. 7, 1–9. https://doi.org/10.1038/ncomms11042 Hausmann, A., Toivonen, T., Heikinheimo, V., Tenkanen, H., 2017a. Social media reveal that charismatic species are not the main attractor of ecotourists to sub-Saharan protected areas. Sci. Rep. 1–9. Hausmann, A., Toivonen, T., Slotow, R., Tenkanen, H., Moilanen, A., Heikinheimo, V., Minin, E. Di, 2017b. Social media data can be used to understand tourists’ preferences for nature- based experiences in protected areas. Conserv. Lett. 1–10. Institute for Economics & Peace, 2019a. Global Peace Index 2019: Measuring Peace in a Complex World. Sydney. Institute for Economics & Peace, 2019b. Global Terrorism Index: Measuring the Impact of Terrorism. Sydney. IUCN, 2019. The IUCN Red List of Threatened Species, Version 2019-3 [WWW Document]. URL http://www.iucnredlist.org (accessed 1.20.20). Jonas, A., Mansfeld, Y., Paz, S., Potasman, I., 2010. Determinants of health risk perception among low-risk-taking tourists traveling to developing countries. J. Travel Res. 50, 87–99. https://doi.org/10.1177/0047287509355323

376 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Kaufmann, D., Kraay, A., Mastruzzi, M., 2011. The worldwide governance indicators: Methodology and analytical issues. Hague J. Rule Law 3, 220–246. https://doi.org/10.1017/S1876404511200046 Kay Smith, M., Diekmann, A., 2017. Tourism and wellbeing. Ann. Tour. Res. 66, 1–13. https://doi.org/https://doi.org/10.1016/j.annals.2017.05.006 Levin, N., Ali, S., Crandall, D., Kark, S., 2019. World Heritage in danger: Big data and remote sensing can help protect sites in conflict zones. Glob. Environ. Chang. 55, 97–104. https://doi.org/10.1016/j.gloenvcha.2019.02.001 Lin, D., Hanscom, L., Martindill, J., Borucke, M., Cohen, L., Galli, A., Lazarus, E., Zokai, G., Iha, K., Wackernagel, M., 2018. Working Guidebook to the National Footprint Accounts. Oakland. Moilanen, A., Anderson, B.J., Eigenbrod, F., Heinemeyer, A., Roy, D.B., Gillings, S., Armsworth, P.R., Gaston, K.J., Thomas, C.D., 2011. Balancing alternative land uses in conservation prioritization. Ecol. Appl. 21, 1419–1426. https://doi.org/10.1890/10-1865.1 Moilanen, A., Pouzols, F.M., Meller, L., Veach, V., Arponen, A., Leppänen, J., Kujala, H., 2014. Zonation - Spatial Conservation Planning Methods and Software Version 4 User Manual. Univ. Helsinki. Naidoo, R., Balmford, A., Ferraro, P.J., Polasky, S., Ricketts, T.H., Rouget, M., 2006. Integrating economic costs into conservation planning. Trends Ecol. Evol. 21, 681–687. https://doi.org/10.1016/j.tree.2006.10.003 Naidoo, R., Gerkey, D., Hole, D., Pfaff, A., Ellis, A.M., Golden, C.D., Herrera, D., Johnson, K., Mulligan, M., Ricketts, T.H., Fisher, B., 2019. Evaluating the impacts of protected areas on human well-being across the developing world. Sci. Adv. 5, 1–8. https://doi.org/10.1126/sciadv.aav3006 OPHI, UNDP, 2019. Global Multidimensional Poverty Index 2019: Illuminating Inequalities. Passport Index, 2019. The Passport Index [WWW Document]. URL https://www.passportindex.org/ (accessed 1.15.20). Powers, R.P., Jetz, W., 2019. Global habitat loss and extinction risk of terrestrial vertebrates under future land-use-change scenarios. Nat. Clim. Chang. 9, 323–329. https://doi.org/10.1038/s41558-019-0406-z Rondinini, C., Wilson, K.A., Boitani, L., Grantham, H., Possingham, H.P., 2006. Tradeoffs of different types of species occurrence data for use in systematic conservation planning. Ecol. Lett. 9, 1136–1145. https://doi.org/10.1111/j.1461-0248.2006.00970.x Santarém, F., Pereira, P., Saarinen, J., Brito, J.C., 2019. New method to identify and map flagship fleets for promoting conservation and ecotourism. Biol. Conserv. 229. FCUP 377 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

https://doi.org/10.1016/j.biocon.2018.10.017 Santarém, F., Saarinen, J., Brito, J.C., 2020a. Mapping and analysing cultural ecosystem services in conflict areas. Ecol. Indic. 110, 105943. https://doi.org/10.1016/j.ecolind.2019.105943 Santarém, F., Saarinen, J., Brito, J.C., 2020b. Desert Conservation and Management: Ecotourism, in: Goldstein, M.I., DellaSala, D.A.B.T.-E. of the W.B. (Eds.), Encyclopedia of the World’s Biomes. Elsevier, Oxford, pp. 259–273. https://doi.org/https://doi.org/10.1016/B978-0-12-409548-9.11827-5 Taylor, N.T., Davis, K.M., Abad, H., McClung, M.R., Moran, M.D., 2017. Ecosystem services of the Big Bend region of the Chihuahuan Desert. Ecosyst. Serv. 27, 48–57. https://doi.org/10.1016/j.ecoser.2017.07.017 UNDP, 2019. Human Development Report 2019: Beyond Income, Beyond Averages, Beyond Today - Inequalities in Human Development in the 21st Century, Human Development Report. United Nations Development Programme, New York, USA. UNEP, 2006a. Global Deserts Outlook. United Nations Environment Programme, Nairobi. UNEP, 2006b. Tourism and Deserts: A Practical Guide to Managing the Social and Environmental Impacts in the Desert Recreation Sector. United Nations Environment Programme, Paris, France. UNWTO, 2019. Yearbook of Tourism Statistics, Data 2013-2017, 2019 Edition. Madrid, Spain. Vale, C.G., Pimm, S.L., Brito, J.C., 2015. Overlooked mountain rock pools in deserts are critical local hotspots of biodiversity. PLoS One 10, e0118367. https://doi.org/10.1371/journal.pone.0118367 Venter, O., Sanderson, E.W., Magrach, A., Allan, J.R., Beher, J., Jones, K.R., Possingham, H.P., Laurance, W.F., Wood, P., Fekete, B.M., Levy, M.A., Watson, J.E., 2018. Last of the Wild Project, Version 3 (LWP-3): 2009 Human Footprint, 2018 Release. Venter, O., Sanderson, E.W., Magrach, A., Allan, J.R., Beher, J., Jones, K.R., Possingham, H.P., Laurance, W.F., Wood, P., Fekete, B.M., Levy, M.A., Watson, J.E.M., 2016a. Global terrestrial Human Footprint maps for 1993 and 2009. Sci. Data 3, 1–10. https://doi.org/10.1038/sdata.2016.67 Venter, O., Sanderson, E.W., Magrach, A., Allan, J.R., Beher, J., Jones, K.R., Possingham, H.P., Laurance, W.F., Wood, P., Fekete, B.M., Levy, M.A., Watson, J.E.M., 2016b. Sixteen years of change in the global terrestrial human footprint and implications for biodiversity conservation. Nat. Commun. 7, 1–11. https://doi.org/10.1038/ncomms12558 Waldron, A., Miller, D.C., Redding, D., Mooers, A., Kuhn, T.S., Nibbelink, N., Roberts, J.T., Tobias, J.A., Gittleman, J.L., 2017. Reductions in global biodiversity loss predicted from

378 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

conservation spending. Nature 551, 364–367. https://doi.org/10.1038/nature24295 Waldron, A., Mooers, A.O., Miller, D.C., Nibbelink, N., Redding, D., Kuhn, T.S., 2013. Targeting global conservation funding to limit immediate biodiversity declines. Proc. Natl. Acad. Sci. USA 110, 1–5. https://doi.org/10.5061/dryad.p69t1 Watson, J.E.M., Venter, O., Lee, J., Jones, K.R., Robinson, J.G., Possingham, H.P., Allan, J.R., 2018. Protect the last of the wild. Nature 563, 27–30. https://doi.org/10.1038/ d41586-018- 07183-6 Whyte, B., 2008. Visa-free travel privileges: An exploratory geographical analysis. Tour. Geogr. 10, 127–149. https://doi.org/10.1080/14616680801999984 World Bank, 2019. World Bank Open Data [WWW Document]. URL https://www.worldbank.org/ (accessed 1.15.20). World Health Organization, 2019. Global Health Observatory data repository [WWW Document]. URL https://www.who.int/en (accessed 1.28.20).

FCUP 379 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Appendix E

SUPPLEMENTARY INFORMATION FOR ARTICLE VI

380 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Vulnerability of desert ecotourism to climate change: lessons from the Sahara-Sahel

Frederico Santarém1,2, Gonçalves, D. 3, Jarkko Saarinen4,5, José Carlos Brito1,2

1 - CIBIO/InBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos da Universidade do Porto. R. Padre Armando Quintas, 11, 4485-661 Vairão, Portugal 2 - Departamento de Biologia da Faculdade de Ciências da Universidade do Porto. Rua Campo Alegre, 4169-007 Porto, Portugal 3 - CIIMAR, Interdisciplinary Centre of Marine and Environmental Research, University of Porto, Terminal de Cruzeiros do Porto de Leixões, Avenida General Norton de Matos, S/N, 4450-208 Matosinhos, Portugal 4 - Geography Research Unit, University of Oulu, Finland 5 - School of Tourism and Hospitality, University of Johannesburg, Johannesburg, South Africa

List of Supplementary Figures and Tables:

Fig. E1 – Averages of the eight global climate models (GCMs; see Table 4.5) for two Shared Socio-economic Pathways (SSP 126 and SSP 585; Riahi et al., 2017) projected for the future for the three bioclimatic variables used in this study. Bioclimatic data are as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Data for the future were obtained at 10-minute resolution from the WorldClim version 2.1 climate data for the years 2081-2100 (Fick & Hijmans, 2017) at https://www.worldclim.org.

Fig. E2 – Uncertainty (standard deviation) in future averages of the eight global climate models (GCMs; see Table 4.5) for two Shared Socio-economic Pathways (SSP 126 and SSP 585; Riahi et al., 2017) projected for the future for the three bioclimatic variables used in this study. Bioclimatic data are as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Data for the future were obtained at 10-minute resolution from the WorldClim version 2.1 climate data for the years 2081- 2100 (Fick & Hijmans, 2017) at https://www.worldclim.org.

FCUP 381 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table E1 - Number of tourism hotspots in each country that is impacted (Present) or will be impacted (Shared Socio-economic Pathways: SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017) by high temperature (>30ºC in BIO 01; >45ºC in BIO 05) and will experience precipitation shifts (>1000mm in BIO 12). Bioclimatic variables are coded as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic variables were obtained at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org. See country codes in Fig. 4.7.

382 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. E1. Averages of the eight global climate models (GCMs; see Table 4.5) for two Shared Socio-economic Pathways (SSP 126 and SSP 585; Riahi et al., 2017) projected for the future for the three bioclimatic variables used in this study. Bioclimatic data are as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Data for the future were obtained at 10-minute resolution from the WorldClim version 2.1 climate data for the years 2081-2100 (Fick & Hijmans, 2017) at https://www.worldclim.org. . FCUP 383 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Fig. E2. Uncertainty (standard deviation) in future averages of the eight global climate models (GCMs; see Table 4.5) for two Shared Socio-economic Pathways (SSP 126 and SSP 585; Riahi et al., 2017) projected for the future for the three bioclimatic variables used in this study. Bioclimatic data are as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Data for the future were obtained at 10-minute resolution from the WorldClim version 2.1 climate data for the years 2081-2100 (Fick & Hijmans, 2017) at https://www.worldclim.org.

384 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Table E1. Number of tourism hotspots in each country that is impacted (Present) or will be impacted (Shared Socio-economic Pathways: SSP 126 and SSP 585; Kriegler et al., 2017; van Vuuren et al., 2017; Riahi et al., 2017) by high temperature (>30ºC in BIO 01; >45ºC in BIO 05) and will experience precipitation shifts (>1000mm in BIO 12). Bioclimatic variables are coded as follows: BIO 01 - Annual Mean Temperature; BIO 05 - Maximum Temperature of the Warmest Month; BIO 12 - Annual Precipitation. Bioclimatic variables were obtained at 10-minute resolution from WorldClim database (Fick & Hijmans, 2017) at https://www.worldclim.org. See

country codes in Fig. 4.7.

LIB

ERI

NIA SSU

NIG

SEN

BUF SUD

ETH TUN

ALG EGY

CHA

MAL

CAM MAU MOR

Present 4 12 >30ºC SSP126 10 77 5 131 110 3 84 11 72 BIO SSP585 110 10 12 145 38 17 3 9 150 167 10 15 175 20 7 154 01 Present 230 10 12 150 112 19 6 37 146 174 133 15 179 20 7 169 24 <30ºC SSP126 230 12 73 112 14 6 37 19 76 133 12 95 9 7 97 24 SSP585 120 5 74 2 3 28 19 123 4 15 24 Present 2 >45ºC SSP126 36 16 33 4 2 9 1 BIO SSP585 219 10 8 106 47 6 23 149 145 40 14 175 6 2 96 16 05 Present 228 10 12 150 112 19 6 37 150 186 133 15 179 20 7 169 24 <45ºC SSP126 194 10 12 150 112 19 6 37 134 153 129 15 177 20 7 160 23 SSP585 11 4 44 65 13 6 14 1 41 93 1 4 14 5 73 8 Present >1000mm SSP126 2 SSP585 7 32 1 1 7 Present 10 50 4 6 1 2 2 2 6 35 500- 1000mm SSP126 2 11 61 6 6 1 2 5 2 2 7 43 BIO SSP585 8 5 56 9 5 1 1 12 22 6 56 12 Present 1 10 2 57 12 64 43 6 13 71 18 1 64 1 200- 500mm SSP126 1 8 1 54 11 71 41 6 10 75 18 58 1 SSP585 2 40 9 87 35 6 3 73 18 42 Present 229 43 112 3 37 85 141 127 108 70 23 <200mm SSP126 229 33 112 2 37 78 143 127 102 68 23 SSP585 230 22 112 1 37 62 150 127 84 2 64 24 FCUP 385 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

References Fick, S. E., & Hijmans, R. J. (2017). WorldClim 2: new 1-km spatial resolution climate surfaces for global land areas. International Journal of Climatology, 37(12), 4302– 4315. https://doi.org/10.1002/joc.5086 Kriegler, E., Bauer, N., Popp, A., Humpenöder, F., Leimbach, M., Strefler, J., Baumstark, L., Bodirsky, B. L., Hilaire, J., Klein, D., Mouratiadou, I., Weindl, I., Bertram, C., Dietrich, J. P., Luderer, G., Pehl, M., Pietzcker, R., Piontek, F., Lotze-Campen, H., … Edenhofer, O. (2017). Fossil-fueled development (SSP5): An energy and resource intensive scenario for the 21st century. Global Environmental Change, 42, 297–315. https://doi.org/10.1016/j.gloenvcha.2016.05.015 Riahi, K., van Vuuren, D. P., Kriegler, E., Edmonds, J., O’Neill, B. C., Fujimori, S., Bauer, N., Calvin, K., Dellink, R., Fricko, O., Lutz, W., Popp, A., Cuaresma, J. C., KC, S., Leimbach, M., Jiang, L., Kram, T., Rao, S., Emmerling, J., … Tavoni, M. (2017). The Shared Socioeconomic Pathways and their energy, land use, and greenhouse gas emissions implications: An overview. Global Environmental Change, 42, 153– 168. https://doi.org/10.1016/j.gloenvcha.2016.05.009 van Vuuren, D. P., Stehfest, E., Gernaat, D. E. H. J., Doelman, J. C., van den Berg, M., Harmsen, M., de Boer, H. S., Bouwman, L. F., Daioglou, V., Edelenbosch, O. Y., Girod, B., Kram, T., Lassaletta, L., Lucas, P. L., van Meijl, H., Müller, C., van Ruijven, B. J., van der Sluis, S., & Tabeau, A. (2017). Energy, land-use and greenhouse gas emissions trajectories under a green growth paradigm. Global Environmental Change, 42, 237–250. https://doi.org/10.1016/j.gloenvcha.2016.05.008

386 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

Appendix F

LIST OF OTHER PUBLICATIONS

FCUP 387 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

388 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

FCUP 389 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

390 FCUP Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel

FCUP 391 Ecotourism development for biodiversity conservation and local economic development in remote regions: A multi-scale approach in Sahara-Sahel