* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download 4 Modelling impacts of drivers on biodiversity and ecosystem
Effects of global warming on humans wikipedia , lookup
Climate change and poverty wikipedia , lookup
Surveys of scientists' views on climate change wikipedia , lookup
IPCC Fourth Assessment Report wikipedia , lookup
Climate change feedback wikipedia , lookup
Climate change, industry and society wikipedia , lookup
Numerical weather prediction wikipedia , lookup
Global Energy and Water Cycle Experiment wikipedia , lookup
Climate change in Saskatchewan wikipedia , lookup
Hotspot Ecosystem Research and Man's Impact On European Seas wikipedia , lookup
First Order Draft IPBES Deliverable 3c DO NOT CITE 4 Modelling impacts of drivers on biodiversity and ecosystem functioning 5 10 15 20 25 Coordinating Lead Authors: Lluís Brotons, Villy Christensen & N H Ravindranath Lead Authors: Olivier Maury, Vania Proença, Pablo Peri, Mingchang Cao, Baris Salihoglu, Jung Hwa Chun. Contributing Authors: Rajiv kumar Chaturvedi, Nicolas Titeux… This chapter describes how to use models that represent ecosystems and biodiversity and account for the impacts of human activities. The key messages from the synthesis are: 1. Ecosystems are open systems that are far too complex to be understood and predicted without formal representations (i.e. models). 2. A diverse range of drivers impact global biodiversity and influence ecosystem stability. Modelling biodiversity requires the consideration of relevant drivers and ecological processes to generate projections that are credible. 3. Clarifying the role of modelling to assess the impacts of changes in direct drivers on biodiversity and ecosystem processes might assist policy making over a range of scales. 4. Biodiversity models rest on processes and assumptions. These must be clearly and transparently communicated in order to be of use in informing decision-making. 5. Ecosystems are complex in nature and interact in complex ways with direct drivers. Models should balance complexity and simplicity considering both the decision-making context and an explicit prioritization of processes and feedbacks to be explicitly accounted for. 6. While important gaps remain in the link between ecological theory and models, different approaches are available to conduct assessments and contribute to scenarios development. 4.1 Introduction 30 35 40 Biodiversity and ecosystem dynamics are complex, and environmental drivers, both of natural and anthropogenic origin, induce changes that are commonly translated into changes in ecosystem services and human wellbeing. The present chapter focuses on the approaches and methods currently available to explicitly link these changes in the environment with biodiversity responses from community composition and structure to biogeochemistry fluxes and ecosystem function descriptors. The aim is to identify the range of tools available to unravel mechanisms of change and incorporate this knowledge in models allowing the projection of future biodiversity conditions. The chapter first provides a comprehensive introduction to the context in which biodiversity and ecosystem models are to be developed, and stresses how the decision-making processes should be used to guide model building. Secondly, a description is provided of broad modelling types and drivers to be included in the quest towards better biodiversity projections. Thirdly, we continue with a detailed description of available modelling approaches, and explanations of how environmental drivers of both natural and anthropogenic drivers are introduced into the Page 1 of 48 First Order Draft 5 10 IPBES Deliverable 3c DO NOT CITE models, and connected with the respective biological components. Fourthly, we describe the main pros and cons of the different approaches in different contextual applications. Since feedbacks between biodiversity responses and environmental drivers may be important and complexity in biodiversity modelling requires careful integration of the spatial scale and contexts, we illustrate current challenges in these ambits. The issues associated with the sources of uncertainty in model projections are of outmost importance in the context of biodiversity projections for IPBES, and we describe this topic in depth in the context of biodiversity and ecosystem modelling. We finally identify the major challenges for biodiversity projections in the context of the IPBES programme, and highlight major avenues for development to make biodiversity projections more useful for policy makers at a range of administrative scales. 4.1.1 Zooming inside the general IPBES scheme. Links to human drivers and ecosystem services 15 20 25 30 35 40 Currently we are in an era that has been called the Anthropocene where human activities have become global geophysical forces, and major drivers of global environmental change (Steffen et al., 2007). According to Rockström et al. (2009), humanity has transcended the “safe operating space” of the planet with respect to biodiversity loss that play an important role as regulator of ecosystem processes. Traditionally, the discipline of ecology studies aspects related to the ecosphere, and the discipline of economics focus on such flows in social systems. However, within the IPBES framework, there is a need of cross-disciplinary fields such as ecological economics and industrial ecology to understand biodiversity and ecosystem services changes. Thus, environmental problems at the global or regional level may reanalyse the perspective of human-environment systems where social and ecological aspects are interacting at multiple temporal and spatial scales (Clark, 2010). In this context, industrial ecology has in the last few decades emerged as a field aiming at a sustainable development of the anthroposphere, which is the interface between the ecosphere and society. Biodiversity plays multiple roles in the conceptual chain linking direct drivers to nature’s benefits to people. Specifically, it may regulate the ecosystem processes that generate final ecosystem services. Indeed, biodiversity itself may constitute a final ecosystem service, or even a good that is directly enjoyed by people (Mace et al. 2012). In the first case, biodiversity attributes affect the development of ecosystem processes (Cardinale et al. 2012), such as nutrient cycling (Handa et al. 2014), primary productivity (Cardinale et al. 2007) or water infiltration (Eldridge and Freudenberger 2005), which in turn give rise to final ecosystem services, including material (e.g., medicinal plants) and non-material (e.g., water regulation) outputs. In the second case, biodiversity elements are themselves material outputs, with direct use value, such as medicinal plants or fish, but which require human capital inputs (e.g., labour, transport) before being enjoyed by society. Finally, biodiversity elements are a good if directly enjoyed by people without any additional input, which is the case of charismatic species and ecosystems. Although biotic and abiotic ecosystem components interact and are both essential to ecosystem functioning the focus of this chapter will be on the biotic components, as framed by the IPBES conceptual framework (Figure 4.1). Page 2 of 48 First Order Draft 5 10 15 IPBES Deliverable 3c DO NOT CITE Figure 4.1. Role of biodiversity and ecosystems in the IPBES conceptual framework. The present chapter concentrates on the impact of drivers of environmental changes (stressing the role of anthropogenic drivers, Chapter 3) on biodiversity and ecosystem processes as determinants of ecosystem goods and services (Chapter 5) The role of biodiversity as a regulator of ecosystem processes or as a material output (either a final service or a good), defines the variables of interest when assessing and projecting the impacts of direct drivers. For instance, community data such as species diversity (Cardinale 2007, Mace et al. 2012) or habitat structure (Eldridge and Freudenberger 2005) may be particularly important to assess the impact of drivers when biodiversity has a regulatory role, while population data, such as species distribution (Gaikwad et al. 2011), or population structure (Berkeley et al. 2004) would be more adequate when biodiversity elements have a direct use value. The acknowledgement of the different roles of biodiversity follows an anthropocentric perspective that has ecosystem services, the material and non-material benefits generated by nature, as its main end. In parallel to utilitarian values biodiversity has its own intrinsic value, which is independent of human demand or appreciation. 4.1.2 Decision making context in biodiversity models 20 25 30 Models allowing the assessment of how environmental drivers of change induce changes in biodiversity or ecosystem functions are essential tools to support decision-making. To be effective, models should be able to capture the problem motivating their use and to answer information needs. A formal and accurate definition of the decision-making context and objective is therefore essential in this process (Guisan et al. 2013). A precise definition of the decision-making context will guide the decision of the modelling framework including model complexity, spatial and temporal scales or response variables and data requirements. Response variables should be sensitive to the pressures underlying alternative management scenarios or addressed by policies, and, if possible, be responsive in temporal and spatial scales that are relevant for policy strategies. For example, small farmland birds are responsive to agro-environmental schemes implemented at the field scale whereas large farmland birds are not, being more affected by the conditions found at the larger spatial scales (Concepcíon and Díaz 2011). Moreover, response variables should also be representative of the biodiversity Page 3 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE attributes underpinning the benefits of nature that are valued in a given decision-making context (see Section 4.1). 5 10 15 Regarding model scope, models should be adjusted to the specific requirements of the decision-making context. Models could rely on observed data to describe the relationship between pressures and response variables, explicitly describe the processes linking those variables, or follow an intermediate approach. The explicit inclusion of mechanisms in modelling approaches will be relevant whenever the understanding of the underlying dynamics is necessary to guide management and where changing environmental conditions call for a mechanistic approach (Gustafson 2013; Collie et al. 2014). The use of correlative approaches, on the other hand, will be adequate if there is limited knowledge about the underlying mechanisms or when model outputs are able to capture the dominant response patterns that are needed to inform policy, such as the evaluation of large scale conservation initiatives (Araújo et al. 2011; Dormann et al. 2012). Regarding model complexity, input data requirements should be balanced against data availability and quality, namely the spatial and temporal resolution of available data, as the lack of or the use of inadequate input data may compromise model feasibility and results quality (Collie et al. 2014). 4.1.3 Why and for what purpose do we need models? 20 25 30 35 40 45 Ecosystems are open systems that are far too complex and multidimensional to be understood without formal abstraction (e.g., models). As humans, we do experience reality but we can only know it through representations. Models are such representations of reality, expressed in scientific terms, most often using mathematical formalisms. Models are tightening together scientific concepts (e.g., energy, population, biomass) using functional hypothesis and the logical grammar of mathematics. They are useful to document the ecological world scientifically, communicate it, and discuss it. Beyond, models are necessary to conceive and think complex scientific objects such as ecosystems and biodiversity, just as words and sentences allow saying but are also the raw material of thinking. Models are therefore at the core of ecological sciences, and ecologists build and use them to represent, explore, share, and discuss their understanding. Ecological models can have various levels of formalisation, from purely qualitative to fully mathematical. The model’s formalisation provides the logical consistency that is needed to manipulate our representations, to make idealized thought experiments that are not possible in real ecosystems, or predict the behaviour of the modelled system beyond the range of observed conditions. Biodiversity is an abstract concept that refers to highly complex living objects (genomes, organisms, populations, communities, ecosystems, Earth system) that are often difficult to observe comprehensively. For that reason, models are useful to formalize the contextdependent scope of biodiversity according to objective definitions, and therefore facilitate discussion and communication. They furthermore allow testing the theoretical consistency of hypothesis on ecosystem functioning and biodiversity response to specific drivers, and their compatibility with observations. They are therefore essential for improving our understanding of ecological patterns and biodiversity. Page 4 of 48 First Order Draft 5 10 IPBES Deliverable 3c DO NOT CITE Ecological observations often reflect the interaction of multiple processes and convoluted factors acting at various scales. By formalising the ecological complexity at stake, ecological models allow disentangling the factors and processes involved to interpret the data. Revealing theoretical inconsistencies, models stimulate the development of theory. Highlighting observation gaps, models can also be used to guide data collection and stimulate new field and experimental studies. Finally, through their capacity to integrate the effects of various drivers and major interactive processes, and predict their future dynamics, ecological models provide strong basis to build biodiversity scenarios and projections, according to possible evolutions of drivers. They can therefore assist policy makers to understand cause-effect relationship between drivers and the impacts and design efficient policies. 4.1.4 Drivers of environmental change and associated scales 15 20 25 30 35 40 The world has in recent decades experienced global environmental change due to human impact, and this has encouraged research to study the new challenges that biodiversity is exposed to (Pereira et al. 2010). The assessments of the established links between these drivers and biodiversity responses are central to the IPBES and determined by both changes in the environment and to the ecological and physiological processes contributing to the dynamics of these ecological systems even in environmentally static systems (e.g., adaptation can induce range expansion; Lavergne et al. 2010). Thus responses to environmental changes derived directly or indirectly from human activities may be either related to changes in the environment itself, to the biological processes acting within ecosystems or, more frequently, to the combination of both (de Chazal and Rounsevell 2009; Leung et al. 2012). It is therefore important to distinguish between indirect and direct drivers of change in ecosystems and their services, and how changes in ecosystem services have affected human wellbeing. Although biodiversity and ecosystem services experience change due to natural causes, these anthropogenic indirect drivers increasingly dominate current environmental changes. The human sources of these changes lie in two major groups of production and consumption activities: (i) resource extraction with associated processing, use and disposal, and (ii) habitatimpacting measure, e.g., in form of land-use and land-cover change (Ayres, 1989). These activities cause the main direct drivers of environmental change for which we are interested in assessing the responses of biodiversity and ecosystem services. Specifically, temporal and spatial habitat modification in marine and terrestrial habitats, land degradation, the impact of invasive species, climate changes, pollution and exploitation and use of resources (see Chapter 2) induce changes in biodiversity and ecosystem function that are not trivially related to the magnitude of the environmental change recorded. Human impacts on the global environment are operating at a range of rates and spatial scales. Scaling issues are particularly important to assess impacts on biodiversity and ecosystem services because drivers have different impact at different scales. For example, while climate change is a driver that acts at the global scale, habitat modification has an impact on biodiversity and ecosystems services at regional and local scales. In this context, at least half of the ice-free surface of the Earth has already been substantially altered for a variety of human uses (Kates et al., 1990), the methane content of the troposphere has doubled, and the level of carbon dioxide increased by 25%, since preindustrial times (Graedel et al., 1990). Similarly, Page 5 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE humans impact all parts of the world ocean, with a large fraction being strongly impacted by multiple drivers (Halpern et al. 2008). 5 10 15 20 25 A better understanding of habitat modification is of crucial importance to the study of global environmental change. Land uses (agriculture, livestock grazing, settlement and construction, reserves and protected lands, and timber extraction) have cumulatively transformed land cover at a global scale. The main direct drivers affecting this process are external input, (e.g., fertilizer use, pest control, irrigation), technology adaptation and use, resource exploitation, pollutions, species introduction and habitat changes. For marine systems, the main drivers are fisheries (Halpern et al. 2008), notably through bottom-impacting fishing gears (Watling 2005). The consequences of habitat modification have been significant for many aspects of local, regional, and global environments, including climate, atmospheric composition, biodiversity, soil condition, and water and sediment flows. However, global-scale assessments typically mask critical sub-global variations. Local and regional case studies can provide the spatial and temporal resolution required to identify and account for major variations in cause-to-cover relationships and the consequence on biodiversity. But, single-factor explanations, at the macro or the micro scale, have not proven adequate. Many models assessing the impact of environmental drivers on terrestrial ecosystems and biodiversity elements, including those dealing with climate and trace-gas dynamics, require projections of land-cover change as inputs. In this context, Loreau et al. (2003) highlighted that knowledge of spatial processes across ecosystems is critical to predict the effects of landscape changes on both biodiversity and ecosystem functioning and services. Changes in ecosystem connectivity after fragmentation or other anthropogenic and natural perturbations may substantially alter both species diversity and ecosystem processes at local and regional spatial scales. Both increasing and decreasing connectivity can either increase or decrease species diversity and the average magnitude and temporal variability of ecosystem processes, depending on the initial level of connectivity and the dispersal abilities of the organism considered. 30 4.2 Definition of main elements determining ecosystem and biodiversity dynamics 35 40 Scientists and stakeholders in decision-making processes are always faced with the challenge of simplifying reality and selecting key processes and drivers most relevant for their study object (Guisan et al. 2013). Decisions in how and what to include explicitly and what can be simplified or ignored are crucial. The prediction of ecological responses to environmental changes should start with the specification of the major model conceptual components relevant for the case and their critical relationships. In the description of any model of this type, three major conceptual components should be identified: 1. Structural elements describing ecosystem. Target state variable(s) used to describe the biological component of interest, such as biomass, species richness, or habitat heterogeneity (Figure 4.2). State variables should be included based on their sensitivity to pressures and the stability of their response pattern, but also the costs and feasibility of data collection. Page 6 of 48 First Order Draft 2. 5 3. IPBES Deliverable 3c DO NOT CITE Environmental and biotic factors whose spatial dynamics have a direct or indirect effect on the biological component of interest (e.g., climate changes inducing species extinction). In a context of environmental change, changes in the value of environmental and biotic factors will affect the value of state variables. Ecological processes relevant in determining changes in the biological component (e.g., species distribution dynamics such as colonization and extinction). Figure 4.2. Summary of some of the predicted aspects of climate change and some examples of their likely effects on different levels of biodiversity. From Bellard et al. 2012. 10 15 20 Biodiversity models, as other mathematical models in environmental sciences, are made of a set of components, namely, state variables, external variables, mathematical equations, and parameters (Jørgensen 1994; Smith and Smith 2007, Soetaert and Herman 2009). State variables correspond to the biodiversity variables of interest, the ones that describe the state of the system, such as biomass, species richness, or habitat heterogeneity (Figure 4.2). State variables should be included based on their sensitivity to pressures and the stability of their response pattern, but also the costs and feasibility of data collection (Dale and Beyeler 2001). External variables (or ‘drivers’) correspond to the external factors that influence the state of the system, such as climate or nutrient inputs. In a context of environmental change, changes in the value of external variables will affect the value of state variables. In biodiversity models, Page 7 of 48 First Order Draft 5 10 IPBES Deliverable 3c DO NOT CITE mathematical equations establish the link between environmental factors (or external variables) and state variables by describing the relationship between them. Models may feature equations that explicitly describe the processes controlling that relationship (mechanistic models), or describe the shape of the response curve by fitting observed data to a predefined function, which is assumed to represent well that process or use a pure statistical approach to derive a relationship from measured values. Parameters are constants in the mathematical equations, which are specific for the system being modelled. The links between external and state variables, and between state variables, as well as the type of processes involved, should be identified at the early stages of the modelling process, namely through the development of a conceptual model (Jørgensen 1994, Smith and Smith 2007). The conceptual model, which can take the form of a diagram, helps to clarify the links between variables but also to define model assumptions – including space and time boundaries (Jørgensen 1994). 15 20 25 30 35 The nature of driver impact on the biological process is key in determining the nature of the model choice and the inclusion of multidisciplinary expertise in the model building process (Guisan et al. 2013). In complex approaches, the impact of drivers on the biodiversity component is compacted through correlative approaches (e.g., SDMs) making the model building process very easy. However, this kind of shortcuts should be performed with caution and full recognition of the assumptions associated to the decision (e.g., ignoring indirect pathways and feedback between element). In a context of environmental change, the effect of environmental pressures on state variables can be direct, (e.g., loss of tree cover after deforestation) or mediated by biological processes, (e.g., ocean acidification and warming affecting coral recruitment and growth, and hence coral abundance and reef structure). In addition, processes also mediate interactions among state variables, (e.g., trophic cascades). The choice of the state variables and processes to figure in a model requires some previous knowledge about the system dynamics. The more realistic a model is, the more complex it will be as realism tends to require description based on solid knowledge of the multiple ecosystem components and processes involved. There is a trade-off with regard to a models predictive capability involved here, where increased complexity leads to a decrease in predictive capability (Walters 1986). Because biodiversity response to environmental change can assume many forms as a consequence of its inherent complexity, there is a large diversity of potential indicators or state variables. A way of addressing this diversity is to reduce it to a few meaningful dimensions. 4.2.1 Describing biodiversity and ecosystem functioning 40 45 Biodiversity indicators or state variables can be arranged along two dimensions representing key aspects of biodiversity complexity, that is, biological organization (species, populations, ecosystems, etc.) and biodiversity attributes (composition, structure, and function). These two dimensions define a conceptual space that can be useful for the identification of state variables and processes ( Table 4.1). More specifically, state variables correspond to composition and structural elements, such as species richness or biomass, and processes to function elements, such as primary productivity, herbivory or competition. Composition and structure emerge from Page 8 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE processes but also affect them (Dale & Beyeler 2001). From an ecological perspective, composition and structure variables describe the structural elements of ecosystems, while processes describe the fluxes of energy and matter and the interactions within and between organization levels. 5 Table 4.1 Examples of compositional, structural, and functional biodiversity variables (Noss 1990, Dale and Beyeler 2001). Based on the representation of the key characteristics of composition, structure and function (from Dale & Beyeler 2001, based on Franklin 1988 at represent the many aspects of biodiversity that warrant attention in environmental monitoring and assessment programs. Individuals Species/ Populations Community/ Ecosystem Landscape/ Region Composition Structure Function Genes Presence, abundance, cover, biomass, density Species richness, evenness and diversity, similarity Genetic structure Population structure, range, morphological variability Canopy structure, habitat structure, Habitat richness Spatial heterogeneity, fragmentation, connectivity Genetic process Demography, dispersion, growth, metabolism, phenology Species interactions (herbivory, predation, competition, parasitism) Succession, decomposition Landscapes processes (hydrologic processes, geomorphic processes), disturbances 10 15 20 25 30 Ecosystems are open systems. They harness solar energy and transfer it through their various structural elements and organization levels, via different biological and ecological processes. At the biosphere level, water and nutrients, such as carbon, nitrogen and phosphorus, are key structural elements of all living components, and key abiotic components of ecosystems. Their flux across the Earth system is described by the biogeochemical cycles. This flux of energy permits life on Earth and fuels the ecological functions that are useful for societies, (e.g., ecological services). To model the dynamics of biodiversity, the major ecological processes involved in the transfer of energy through ecosystems must be taken into account. These include production, consumption, respiration, and recycling. Other processes such as regulation and evolution are critical to the maintenance of ecosystems over time and also involve biodiversity. Broadly speaking, mechanisms determining ecosystem dynamics can be divided into three broad groups (Lavergne et al. 2010). First, we find mechanisms related to the ecology of the species populations, such as habitat selection, dispersal, and population dynamics. Those processes are primarily determined by species traits expressing the capability of the target species to deal with environmental variability in space and time, (e.g., Hanski et al. 2013; Thuiller et al. 2013). At this level of biological organization, the main structural elements, besides individuals, are the cohorts or life stages, such as seedlings and mature trees. Primary production and respiration are major ecological processes, occurring at the organism level and underlying population dynamics. Organic matter from primary production is at the basis of all Page 9 of 48 First Order Draft 5 10 15 20 25 30 35 40 IPBES Deliverable 3c DO NOT CITE life on Earth. Numerous factors such as light, the availability of inorganic nutrients, water and temperature influence primary production and have to be considered explicitly in models. Respiration, which comprehends all the living processes using oxygen, is at the core of metabolism. Respiration can be considered at every organization levels, from mitochondria to ecosystems. At the organism level, respiration processes are influenced by many factors including the species considered (body-size scaling rules imply that many metabolic processes vary with the maximum size that a species can reach, Kearney et al. 2010a), the size of individuals, their condition, the availability of food, oxygen and temperature. At the population level, respiration integrates the metabolism of all individuals. It is therefore highly dependent of the size- and state-structure of the population. At the community level, respiration integrates the metabolism of all populations and is therefore controlled by their relative abundance and the structure of the community. Modelling the variability and environmental influences on metabolic biomass production and dissipation therefore requires integrating over several organisation levels, from individuals to communities (Maury and Poggiale, 2013). In the second place, species interactions can restrict or expand the set of places that the species is able to inhabit (Davis et al. 1998). Competition, facilitation or trophic relationships are site and species-specific and account for a great deal of variability in the species capability to survive in a given environment. At the community level, the main structural elements of interest are related with the trophic dynamics (e.g., producers, consumers, detritivores) and other species interactions (e.g., plant-pollinator interactions). Consumption and recycling are main processes associated to trophic interactions. Consumption constitutes a major process of ecosystem dynamics that transfers solar energy along food chains, from primary producers up to top predators. Trophic interactions are influenced by various factors, including by spatialtemporal co-occurrence of grazers/predators and their food/prey, which is often constrained by environmental features. In the oceans, both horizontal and vertical distributions and movements of organisms control trophic interactions, but the vertical distributions are especially dynamic and can change over short time scales. Alternative modelling approaches considering explicit modelling of vertical distributions through three-dimensional models with the complexity this entails have to be explored along with implicit modelling of vertical distributions. Recycling is a major component of the dynamics of ecosystems that must considered in models for closing the system-level mass-balance. Recycling has a critical importance in the biogeochemistry of the Earth System. It is in particular responsible for the storage of carbon in soils and its export to the deep marine sediments via the biological carbon pump (Sarmiento and Gruber, 2006). Finally, evolutionary processes including contemporary microevolution and macro evolutionary trends, allow for critical changes in species traits (Lavergne et al. 2010). Any biological process of interest should have an explicit link with the components formulated in the model. However, this link does not need to be one-to-one. Page 10 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE 4.3 Modelling approaches for assessing the impact of drivers on biodiversity and ecosystem functioning 5 10 15 20 25 30 35 Various types of ecological models are used for assessing the impacts of large-scale environmental drivers on the ecosystem. Ecological models can be classified in two broad categories: phenomenological and mechanistic. Amongst the phenomenological approaches, purely statistical models, that only rest on correlations between variables, must be distinguished from phenomenological process-based models that explicit some relevant processes in a phenomenological way. Mechanistic approaches attempt to derive ecosystem dynamics from first principles, essentially like theoretical physics attempts to describe the nonliving universe. Practically, since the development of a purely mechanistic theory of ecosystems is still to come, existing approaches can be classified along a gradient from purely phenomenological to purely mechanistic. Correlative models use statistical methods to establish relationships between environmental variables and biodiversity data, such as species richness, abundance or distribution (Morin and Lechowicz 2008). These models deliver information on biodiversity patterns and responses to drivers based on empirical observations, but do not explain the mechanisms underlying those patterns and responses (Rahbeck et al. 2007). Model performance and realism is therefore dependent on a careful choice of environmental predictors and of the spatial resolution and extent of the data, but also on the judicious use of the model given model assumptions (Elith and Leathwick 2009, Araújo and Peterson 2012). Figure 4.3. A simple schematic showing the relationship between two observations of a species distribution in the ‘real world’, ‘statistical models’ and ‘dynamic, process based models’. From McInerny & Etienne 2012. Process-based phenomenological models are empirical constructions. They use the inductive scientific method, going from empirical observations to likely generalizations. Unlike purely statistical models, process-based phenomenological models explicitly consider some relevant processes that they represent using mathematics as a tank of flexible (potentially non parametric) shapes that are fitted to data to mimic the observations. That is providing them with some functional capacity and the possibility to interpret the data within a formal theoretical framework. However, process-based phenomenological models may have limited cognitive capacity (they describe rather than they explain) and they are limited by (1) the sensitivity of the system dynamics to the arbitrariness of the mathematical form used to represent the process, (2) the sensitivity to the data used to estimate the parameters and the Page 11 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE impossibility to predict beyond the range of observed conditions, (3) the disconnection of processes that leads to a lack of internal consistency and over-parameterization. 5 10 15 20 25 At the other end of the formalization gradient, pure mechanistic models also termed as theoretical models are axiomatic constructions (Gallien et al. 2010). Like in theoretical physics, they use the hypothetico-deductive scientific method, starting from hypothesis (the axioms) to deduce predictions that can be tested empirically, to falsify or conversely to corroborate the hypothesis made (but never to prove it or “validate” it). Mechanistic models are based on few basic rules and general principles in which processes are explicitly accounted for and from which their dynamics can be deduced logically and mathematically. They consist in mathematically derived parameterizations where parameters are meaningful constants. They are based on clear processes and causality chains formulated from “first principles”. Mechanistic models of biodiversity have to be internally consistent with both biological theory and mathematical rules. They must represent the biological and ecological processes, respecting dimension rules and verifying the principle of energy and mass conservation. In the literature, the term mechanistic is often used in a wider perspective to include models explicitly accounting for processes in their internal structure (i.e. correlative mechanistic models, Kearney et al. 2010). In this chapter we use the term process based in the context of explicit implementation of ecological processes in the model and mechanistic to describe models with stronger foundations on ecological theory. Mechanistic models of biodiversity may help improve our understanding, interpret observations, guide data collection, and stimulate new field and experimental studies. Mechanistic models have a high cognitive capacity (they explain more than they describe). However, they are often very complex, empirically undetermined, and difficult to use practically. 4.3.1 Phenomenological correlative modelling approaches 30 35 40 45 Phenomenological correlative models are frequently used to assess the impacts of human activities on biodiversity, forecast future impacts of environmental changes, support human productive activities (e.g., enhance agricultural production) and conservation actions (e.g., identify sites for translocations and reintroductions, predict the location of rare and endangered species), understand species’ ecological requirements, among other uses (Peterson, 2006; Muñoz et al. 2011). Correlative models have the advantage of offering greater interpretability as to causation of phenomena, and permit predictability of phenomena that depend on the differences between components—e.g., the invasive potential of a species depends on the difference between potential and actual distributional areas (Peterson, 2006). Correlative models can be applied at all spatial scales. The choice of the appropriate spatial scale should consider the scale of variation of relevant environmental predictors (Elith and Leathwick 2009). For instance, the effect of climate variables is better assessed at large spatial extents, such as regions, and coarse resolution data may be acceptable, whereas the effect of soil nutrients requires fine resolution data, to cover fine-scale variations, and is usually modelled at smaller extents, such as landscapes. When the selected environmental predictors act at different scales, hierarchical models with nested sub-models can be used (Elith and Page 12 of 48 First Order Draft 5 10 15 20 IPBES Deliverable 3c DO NOT CITE Leathwick 2009). Similarly, advances in modelling marine ecosystems will require coupling of different trophic levels that may have different resolution. A way in representing biodiversity in complex marine systems would be to concentrate the detail of representation at the target species. Regarding temporal scales, correlative models are often static (i.e., assume a constant time frame) failing to capture species or community dynamics, such as species dispersal. Nevertheless, temporal predictors, such as variability of food resources, may be added to models to capture variation in the state of biodiversity variables. More complex dynamics, such as species dispersal, can be integrated in modelling approaches, such as Dynamic Bioclimate Envelope models (DBEM), where an envelope niche model is combined with dynamics of species dispersal (Cheung et al. 2008). An important caveat of correlative models, concerns extrapolation to new spatial and time frames, namely hind-casting and forecasting applications, since the conditions associated to training data (i.e., the data used to fit the model), may not be constant over time (Elith and Leathwick 2009, Araújo and Peterson 2012) or be inadequate to represent the conditions found outside their area of distribution. For instance, the data informing niche models are not drawn from the entire abiotic niche or even from the potential distribution, but from the actual area of distribution, which calls for caution when interpreting results (Araújo et al. 2005; Pearson and Dawson 2003). Alternative approaches based explicitly on mechanisms are likely to be robust under novel environmental combinations but are limited by the availability of data to build the models. In contrast, data required to fit correlative models is readily available across a range of scales and the models can implicitly capture many complex ecological responses. Because of this, Elith et al. (2010) anticipate ongoing use of correlative models for biodiversity projections. 25 30 35 Species occurrence and abundance are common state variables in correlative modelling approaches at the species level whereas variables of community diversity and structure are common in community-level approaches (Iverson et al. 2001). Environmental datasets often comprise climatic variables (e.g., mean annual temperature, annual precipitation, and annual solar radiation), topographic variables (e.g., average elevation, elevation coefficient of variation, elevation range), and habitat variables (e.g., habitat type, habitat extent). For instance, a study to assess the impacts of climate change on the distribution of Pinus densiflora in South Korea (Chun and Lee 2013) utilised 27 environmental variables from 4 categories of climate characteristics, geographic and topographic characteristics, soil and geological properties, and MODIS EVI, and about 4,000 species occurrence and abundance records from National Forest Inventory sites across the whole country. 4.3.1.1 Species-level modelling 40 Species distribution models (SDMs), also known as niche models or bioclimatic envelope models, have been widely used to model the effects of environmental changes on species distribution across all realms (Pearson & Dawson 2003, Elith and Leathwick 2009). A bioclimatic envelope can be defined as a set of physical and biological conditions that are suitable to a given species, it is generally identified through the relationship between species observed distribution and environmental attributes using statistical methods (e.g., generalized Page 13 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE additive model, Luoto et al. 2005) or artificial intelligence models (e.g., artificial neural network, Pearson et al. 2002). 5 10 15 20 25 30 35 40 Main critiques to bioclimatic envelope models relate to their omission of biotic interactions, evolutionary change and species dispersal processes (Pearson & Dawson, 2003). Still, examples exist where bioclimatic envelope models were developed to include some of these processes. The Dynamic Bioclimate Envelope Model developed by Cheung et al. (2008) simulates changes in the relative abundance of marine species through changes in population growth, mortality, larval dispersal and adult movement following the shifting of the bioclimate envelopes induced by changes of climatic variables. The model does not account for species interactions and potential food web changes. A new dynamic bioclimate envelope model is being developed to account for effects of ocean biogeochemistry such as oxygen level and pH on the ecophysiology and distribution of marine fish (Stock et al. 2011). SDMs often use species presence-absence data, however, in some situations there is only access to presence data (e.g., museum data). In those situations other models, such as GARP and MaxEnt, can be used (Elith and Leathwick 2009). GARP (Genetic Algorithm for Rule set Production) is a machine-learning system that has shown predictive ability in delineating species’ ecological niches and predicting geographic distributions (Peterson et al., 2002). At an initial stage, GARP produces four types of rule sets (atomic, range, negated range, and logistic regression) explaining the relationship between the distribution of species presence data and explanatory environmental variables. Then it uses an iterative process of rule modification and selection until it reaches the most appropriate set of rules, which can then be used to predict species distributions. MaxEnt is a presence/background method, that is, model fitting uses presence-only data and data on environmental variation across the study area (i.e., background data) (Peterson et al. 2011). The algorithm estimates the species range according to a probability distribution of maximum entropy (i.e., closest to uniform) constrained by information on environmental data at presence sites (Araújo et al. 2007, Phillips and Dudík 2008, Merow et al. 2013). Maxent has been used for several purposes, such as, to assist conservation planning, to predict potential distributions for invasive species, and to predict climate and land use change impacts (Elith et al. 2010). Case study: Conservation of Iberian lynx populations (Fordham et al. 2013) Decision-making context: Iberian lynx populations have suffered severe declines during the last century due to habitat change, changes in abundance of preys (due to habitat change and overexploitation), direct persecution and road killing. Today, the Iberian lynx is the world’s most endangered cat. On-going conservation actions include habitat protection and restoration for the species and its preys (mainly rabbits), ex-situ breeding and reintroductions. Policymakers are considering the establishment of new Iberian lynx populations in areas within its recent historical range, namely in different autonomous regions in the southern half of Spain. However, this solution may be compromised by climate change due to impacts on habitat suitability and barriers to dispersal. An alternative would be to perform reintroductions in habitat-favorable areas across all Iberian Peninsula, even if outside the recent historical range. Page 14 of 48 First Order Draft 5 10 15 20 25 IPBES Deliverable 3c DO NOT CITE Modelling approach: The future distribution of Iberian lynx populations was assessed using coupled ecological niche models (ENM) of Iberian lynx and European rabbit range dynamics under different climate change and conservation scenarios. Climate change scenarios include a high-CO2 stabilizing reference scenario and a greenhouse gas mitigation Policy scenario. Conservation scenarios, included a business as usual scenario, an extramanagement scenario (no reintroductions), and two reintroductions scenarios, within the recent historical range and Peninsula-wide. ENM were coupled to metapopulation simulations of source-sink dynamics to account for stochastic changes in Iberian lynx populations driven by changes in prey abundance, due to disease, climate and land-use changes, and by changes in habitat suitability, due to climate and landuse changes. Main results: Probability of extinction is high for the business as usual and extra management conservation scenarios (>85%) for the two climate-change scenarios. Planned reintroduction programs, on the other hand, can reduce extinction risk (<5%). Planning reintroductions in habitat-favorable areas across all Iberian Peninsula may be more successful than the restricting reintroductions to the recent historical range. 4.3.1.2 Species traits approaches 30 Trait based ecological risk assessments (TERA) is a trait based approach to assess ecological responses to natural and anthropogenic stressors based on species characteristics related to their functional roles in ecosystems (Baird et al. 2008). In case there is enough information on functional characteristics and roles of taxonomic species or communities in the ecosystems, TERA has potentials in ecological risk assessment adding trait attributes to map onto existing taxonomic entities but more cooperative effort is needed to address the challenges such as task of trait definition, linkage of traits to specific environmental drivers, and the extraction of trait information by observation, experiment, and literature data mining (Baird et al. 2007). 35 4.3.1.3 Community-level modelling 40 Community-level models are used to produce information at the collective level (Ferrier, 2002). The community metric of interest can take a number of forms, such as predictive mapping of community types (groups of locations with similar species composition), species groups (groups of species with similar distributions), axes or gradients of compositional variation, levels of compositional dissimilarity between pairs of locations, and various macroecological properties (e.g., species richness) and even phylogenetic diversity (Ferrier & Guisan, 2006). Community-level modelling offers an opportunity to move beyond species-level predictions to predicting broader impacts of environmental changes (e.g., Hilbert & Ostendorf Page 15 of 48 First Order Draft 5 10 15 IPBES Deliverable 3c DO NOT CITE 2001; Peppler-Lisbach & Schröder 2004), which may be relevant in certain decision-making contexts. Examples are when time and financial resources are limited, when existing data is spatially sparse or the knowledge on individual species distribution is limited (Ferrier, 2002), and when species diversity is beyond what can feasibly be modelled at the individual species level. Three community-level modelling approaches have been proposed (Ferrier & Guisan, 2006): (1) ‘Assemble first, predict later’, whereby species distributions are first combined with classification or ordination methods and the resulting assemblages are then modelled using machine learning or regression-based approaches, (2) ‘Predict first, assemble later’, whereby individual species distributions are modelled first and the resulting potential species distributions are then combined (i.e. the result is in fact the summation of individualistic models), and (3) ‘Assemble and predict together’, whereby species distribution models are fitted using both environmental predictors and information on species co-occurrence. Reviews on usefulness of community-level modelling can be found in Ferrier et al. (2002) and Ferrier & Guisan (2006) Table 4.2. The six main type of spatial output that can be generated using community level modeling (Ferrier and Guisan, 2006) Community metrics Description Structure of derived grid layers Individual species Predicted distribution of multiple species, as for species-level modelling Community types Each’ community type’ defined as a group of locations (grid cells) that closely resemble one another in terms of predicted species composition. Grouping normally achieved through some form of numerical classification Species groups Each ‘species group’ defined as a subset of species with similar predicted distributions. Grouping again achieved through numerical classification, but in this case the objects classified are species rather than locations A set of continuous axes (or gradients) representing dimensions of a reduced space that summarizes the compositional pattern exhibited by multiple species. These axes most commonly derived through some form of ordination The predicted level of dissimilarity in community composition between all possible pairs of grid cells in a region A separate layer for each species, indicating the predicted probability of occurrence or abundance of that species in each cell Either (i) a single layer with each cell assigned exclusively to one community type (depicted as a map with different colours indicating different types) or (ii) a separate layer for each community type, indicating the probability of that type occurring in each cell (depicted as multiple grey-scale or colour-ramp maps) A separate layer for each species group, indicating the predicted prevalence or abundance of that group in each cell (depicted as multiple grey-scale of colour-ramp maps) Axes of compositional variation Levels of compositional dissimilarity between pairs of cells Macroecological properties Most commonly modelled property is species richness, either of a whole group (e.g. all vascular plants) or of a functional subgroup (e.g. annuals and trees). Many other macro-ecological properties (e.g. mean range size and endemism) can potentially be modelled A separate layer for each axis, indicating the predicted score for that axis in each cell (depicted either as multiple grey-scale or colour-ramp maps, or as a single map by assigning each of the first three axes to a different colour dimension, e.g. red, blue, green) In theory a complete matrix of pair-wise dissimilarities, but in practice these values are usually predicted dynamically as required by the application of interest (difficult to depict spatially without prior conversion to community types or axes of compositional variation) A separate layer for each macro-ecological property (depicted as a grey-scale or colour-ramp map) Page 16 of 48 First Order Draft 5 10 15 20 25 30 35 40 IPBES Deliverable 3c DO NOT CITE Also important at the community-level are species-area relationship (SAR) models, which use habitat area as a predictor of community species richness. SARs have been applied to a wide range of taxa and across all scales, from local to global. SAR functions are often used to predict the impacts of changes in habitat availability, driven by land use change (e.g., van Vuuren et al. 2006, Desrochers et al. 2011) or climate change (e.g., Malcom et al. 2006, van Vuuren et al. 2006), on community richness, but also to assess the impacts of direct exploitation on community parameters, such as species-turnover rates (e.g., Tittensor et al. 2007). Reviews on the use of SARs can be found in Rosenzweig 1995, Drakare et al. 2006, and Triantis et al. 2012). 4.3.2 Process-based and mechanistic models Considering the relevant processes provides process-based phenomenological models with some functional capacity and the possibility to interpret the data within a formal theoretical framework. However, process-based phenomenological models have poor cognitive capacity (they describe rather than they explain) and they are limited by (1) the sensitivity of the system dynamics to the arbitrariness of the mathematical form used to represent the process, (2) the sensitivity to the data used to estimate the parameters and the impossibility to predict beyond the range of observed conditions, (3) the disconnection of processes that leads to a lack of internal consistency and over-parameterization. Let’s look at a few illustrative examples to clarify these limitations: 1. Even though they can theoretically be derived from ecological mechanisms, functional response curves are usually used as flexible and practical regression functions linking the amount of food eaten by a predator to the density of prey, disregarding the underlying mechanisms that control their shape. Yet, when embedded in ecosystem models, the choice of the functional response’s form (e.g. type I, II or III) leads to profound differences in the modelled ecosystem dynamics (Fulton, 2003; Anderson et al., 2010). 2. Other classical examples of phenomenological representation include the allometric equations that are typically used for representing the various processes of metabolism (e.g., respiration is proportional to some power – named the allometric coefficient – of individual weight). Because allometric functions are stationary and do not explicit the influence of environmental factors on metabolism, any change of environmental conditions is absorbed by a corresponding change in parameter value. Changing the dataset used to estimate the allometric coefficients therefore leads to different estimated values and associated responses of metabolism to weight (e.g. Bonnet et al., 2003). As a consequence, parameters are meaningless system- and state-specific “constants that vary” in space and time to capture the observations. 3. Because metabolism includes very different processes with specific dynamics, fitting allometric regressions to different processes of metabolism (e.g., respiration, reproductive output, age at maturity, aging mortality) leads to as many allometric coefficients as processes considered. Using independent allometric curves to represent the various metabolic processes at stake in an ecosystem model therefore leads to overparameterization and do not consider the interactions between the processes represented. Page 17 of 48 First Order Draft 5 IPBES Deliverable 3c DO NOT CITE Process-based models allow a more explicit representation than correlative approaches of ecological processes mediating biodiversity responses to environmental drivers. However, and despite their wide use in biology and ecology, process-based phenomenological ecosystem models, as all model developments suffer from fundamental and practical limitations that should be explicitly acknowledged. Examples Various strategies and levels of process formalisation can be distinguished amongst the available process-based phenomenological models of ecosystem: 10 15 20 1. Box models. This is the simplest and most developed category. It describes the ecosystem structure as a set of bulk functional groups that are connected together by fixed uptake/predation terms based on given functional responses, and that dissipate energy through simple respiration terms that do not explicit the various components of metabolism. The main advantage of these models is that they balance the mass and energy fluxes at the scale of the system represented. The counterpart is that they neglect important phenomenon such as the importance of size in controlling metabolism, predator–prey interactions and life-history omnivory (i.e. diet changes when organisms grow). Examples include nutrient-phytoplankton-zooplanktondetritus (NPZD) type models (e.g. PISCES ref, MEDUSA ref, ERSEM ref, TOPAZ ref, etc…) that represent 3D marine biogeochemistry in climate models, mass-balanced trophic models (e.g. EwE Polovina, 1984; Christensen and Pauly, 1992; Christensen and Walters 2004) that represent marine communities in 0-D (Ecopath, Ecosim) or in 2-D (Ecospace), often with a focus on upper trophic level organisms. 25 30 35 40 45 2. Age/stage-structured models. These models are box models that are structured along a dimension (usually age or stage) that is supposed to be functionally important. They explicit some processes of metabolism such as growth, reproduction and the agedependence of respiration. This group, among others includes dynamic global vegetation models (DGVM) that represent the dynamics and biogeochemistry of the global vegetation cover, the Atlantis model (Fulton et al. 2005) that represents the trophic interactions and biogeochemistry of marine trophic networks in 3-D boxes, and the dynamic EwE models, which incorporates age-structured population models (Walters et al. 2000; Walters et al. 2010). Spatial population dynamics models (e.g. Lehodey, 2008) use habitat-preference functions to describe population movements with advection-diffusion equations. 3. Size-structured models. Some approaches emphasize the interest of size to structure ecosystem models. In marine ecosystems, size rather than taxonomic identity controls predation (e.g., Shin and Cury, 2004) and metabolism (e.g., Gillooly et al., 2001). Also, size-based models are easier and cheaper to parameterize than food-web models. A wide variety of size-structured models exist. This include 0-D marine ecosystem sizespectrum models, where ecosystem dynamics is projected along the size dimension “from bacteria to whales”, disregarding the metabolic and physiological differences between species (e.g., Platt and Denman, 1978; Benoit and Rochet, 2004; Maury et al., Page 18 of 48 First Order Draft 5 10 15 20 25 30 IPBES Deliverable 3c DO NOT CITE 2007). In the same perspective, Madingley (Harfoot et al., 2014) represents the 2-D energy fluxes through size-structured functional groups with a lagrangian approach, in both terrestrial and marine environments. Recently, the need to consider explicitly the functional role of species diversity in the dynamics of functional groups has emerged. This has led to the development of trait-based approaches (e.g., Andersen and Beyer, 2006; Hartvig et al., 2011), where the community-level dynamics emerges from the interactions of a large number of species that are distinguished solely on the basis of their maximum size, which captures most of the inter-specific differences of metabolism and life history (Kooijman, 1986). Other approaches, like the 2-D lagrangian models, Osmose (Shin and Cury, 2004) and size-based Ecospace model (Walters et al. 2010), represent explicitly a set of key size-structured species and emphasize the importance of size and spatial co-occurrence in controlling trophic interaction and the emergent structure of marine communities. Dynamic Global Vegetation Models (DGVM) The most advanced tool to estimate the impact of climate change on vegetation dynamics at global scale are Dynamic Global Vegetation Models (DGVMs). These are process-based models that simulate various biogeochemical, biogeophysical and hydrological processes such as photosynthesis, heterotrophic respiration, autotrophic respiration, evaporation, transpiration and decomposition. DGVM models simulate shifts in potential vegetation and the associated biogeochemical and hydrological cycles as a response to shifts in climate. DGVMs use time series of climate data and, given the constraints of latitude, topography, and soil characteristics, simulate monthly or daily dynamics of ecosystem processes. DGVMs are most often used to simulate the effects of future climate change on natural vegetation and its carbon and water cycles. DGVMs integrate biogeochemistry, biogeography, and disturbance sub-models, and disturbance is generally limited to wildfires (for example in IBIS wildfires are represented by static values). DGVMs are usually run in a spatially distributed mode, with simulations carried out for thousands of “grids” or geographic points, which are assumed to have homogeneous conditions within each grid. Under DGVM framework, simulations are carried out across a range of spatial scales, from global to landscape. With a growing interest in the biodiversity changes in an ecosystem, these ESM’s generally comprise of the coupling of DGVMs with AOGCMs. Basic structure of a DGVM is given in Figure 4.4. Page 19 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE Figure 4.4. Structure of DGVMs. Source: http://seib-dgvm.com/oview.html 5 10 15 20 25 Dynamic models capture the transient response of vegetation to a changing environment using explicit representation of key ecological processes such as establishment, tree growth, competition, death, nutrient cycling (Shugart 1984, Botkin 1993). Plant functional types (PFTs) are central to DGVMs as, on the one hand, they are assigned different parameterizations with respect to ecosystem processes (e.g., phenology, leaf thickness, minimum stomatal conductance, photosynthetic pathway, allocation, rooting depth), while on the other, the proportion of different PFTs at any point in time and space defines the structural characteristics of the vegetation (Woodward and Cramer 1996). Input and outs: DVMs use spatially explicit time series of climate data (e.g., temperature, precipitation, humidity, sunshine days, wind) and take into account the features of topography and soil characteristics in order to simulate monthly or daily dynamics of ecosystem processes. Plant species are represented as groups with similar physiological and structural properties, termed Plant Functional Types (PFTs), which are designed to represent all major types of plants (Sitch et al. 2008). DGVMs are most often used to simulate the effects of future climate change on natural vegetation and its carbon and water cycles. DGVM provides policy relevant outputs with respect to ecosystems such as shifts in vegetation types, NPP and carbon stock changes, biomes and habitat change. Advantage: Key advantages of using DGVMs include its capacity to simultaneously models the transient responses related to dynamics of plant growth, competition and, in a few cases, migration. As such it allows the identification of future trends in ecosystem function and structure and these models can be used to explore feedbacks between biosphere and atmospheric processes (Bellard et al 2012). 4.3.2.1 Mechanistic models 30 Mechanistic models of biodiversity are tools to improve our understanding, interpret observations, guide data collection, and stimulate new field and experimental studies. Their consistency and structural stationarity (e.g., mechanistic models keep valid in other environments, including beyond the range of observed conditions) provide a basis to build and analyse scenarios and projections. Mechanistic models have a high cognitive capacity (they Page 20 of 48 First Order Draft 5 10 IPBES Deliverable 3c DO NOT CITE explain more than they describe). However, they are often very complex, empirically undetermined and difficult to use practically. Purely mechanistic models and theories in the fields of biodiversity are rare. In practice, they are often incomplete and articulate mechanistic descriptions of the components under focus with phenomenologic parameterizations of the less known or less important phenomenon. However, in recent years there has been growing effort to better place existing approaches into more available ecological theories allowing wider use and greater potential of the models derived from this approach. Some of these examples are the theory based biodiversity model introduced by Thuiller et al. (2013), which used metapopulation theory as the cornerstone for a new generation of biodiversity models. This kind of approaches has the advantage of being explicit about the processes involved and provides interpretable parameters while at the same time has solid foundations in population dynamics. Furthermore, the new developments build on the recent addition of environmental heterogeneity, dispersal limitation, and biotic interactions into the incidence function and include the addition of an evolutionary perspective as a new challenge. 15 20 25 30 35 40 Another theory used in mechanistic modeling is the Dynamic Energy Budget (DEB) theory (e.g., Kooijman 2010) which is a mechanistic theory aiming at capturing the quantitative aspects of metabolism at the organism level from a small set of key assumptions (Sousa et al., 2008a). The DEB theory allows accounting for environmental variability effects on organisms through food and temperature changes. It captures the diversity of all possible living forms on earth in a single mechanistic framework. This allows representing the energetics and major life history traits of all possible species in a community with the same set of unspecific taxa-dependent DEB parameters. The existence of inter-specific scaling rules is a fundamental property of the DEB theory. It emphasizes the fundamental importance of the maximum size of the species and implies that functional biodiversity (in terms of energetics and life history) has actually far less degrees of freedom that one would expect given the high number of species on Earth. This major finding opens the way to a mechanistic understanding of functional biodiversity and ecosystems dynamics. DEB-based individual bioenergetics can indeed be scaled up to build physiologically-structured mechanistic population dynamics models that can in turn be scaled up using interspecific scaling rules to derive size-based mechanistic models of the dynamics of communities (Maury and Poggiale, 2013). The APECOSM model adopts this strategy to represent the dynamics of marine ecosystems in a mechanistic way (Maury, 2010). Embedded in a 3-D mechanistic framework representing organisms movements (Faugeras and Maury, 2007) and environmentally-driven behaviours, the trait-based dynamics of canonical DEB-structured marine communities is modelled from regional to global scales. APECOSM can be coupled to an Ocean General Circulation Models (OGCM) and an Ocean General Biogeochemical Model (OGBM) that provide the timedependent and tri-dimensional environmental forcing variables, both for hind- and forecasting (Lefort et al., 2014; Dueri et al., 2014). 4.3.2.2 Other approaches 45 Deterministic Models: In these models one cause is modeled to have one or many effects. The term “forest dynamics” spans a huge range both in time and space. For example the enzymatic reactions of photosynthesis operate within fractions of a second; foliage development takes a Page 21 of 48 First Order Draft 5 10 15 20 25 30 35 IPBES Deliverable 3c DO NOT CITE few weeks, while tree growth lasts decades to centuries, and the dynamics of soil organic matter span millennia (Bugmann 1994). On the other hand, the germination of a seed takes place on a few square cm3, a sunfleck moving over the forest floor covers a few m2; a dominant tree in the canopy occupies 0.01-0.1 ha, and the quasi-equilibrium of a forest landscape may be reached on the scale of several hectares (Shugart and Urban 1989). There have been two basic approaches to modeling vegetation response to changing climates, static (timeindependent) and dynamic (time-dependent) models. Static or equilibrium models: Static models assume equilibrium conditions in both climate and terrestrial vegetation in order to predict the distribution of potential vegetation by relating the geographic distribution of climatic parameters to the vegetation. The equilibrium approach is implicitly large scale in nature as it ignores dynamic processes. It generally requires far less information and provides estimates of potential magnitude of the vegetation response at regional to global scales. The equilibrium models are restricted to the estimation of the steadystate conditions. Prominent examples of static models include BIOME4 (Kaplan 2003) and Mapped Atmosphere Plant Soil System (MAPPS; Neilson 1995). Biogeochemistry Models: Biogeochemistry models project changes in basic ecosystem processes such as the cycling of carbon, nutrients, and water. These models are designed to predict changes in nutrient cycling and primary productivity. The inputs to these models are temperature, precipitation, solar radiation, soil texture, and atmospheric CO2 concentration. The plant and soil processes simulated are photosynthesis, decomposition, soil nitrogen transformations mediated by microorganisms, evaporation and transpiration. Common outputs from biogeochemistry models are estimates of net primary productivity, evapotranspiration fluxes and the storage of carbon and nitrogen in vegetation and soil. Some of the popular models include BIOME-BGC (Running and Hunt, 1993), CENTURY (Parton et al. 1993), and the Terrestrial Ecosystem Model (TEM; McGuire et al. 1992). Biogeography Models: Biogeography models simulate shifts in the geographic distribution of major plant species and communities. They analyze the essential environmental conditions over entire continents, to estimate the type of vegetation that is most likely to cover a given area. These types of models are best suited for assessing broad-scale changes in vegetation. They are based on ecophysiological constraints, which determine the broad distribution of major categories of woody plants, and response limitations, which determine specific aspects of community composition, such as the competitive balance of trees and grasses. Examples of biogeography models include the Mapped Atmosphere Plant Soil System (MAPSS; Neilson, 1995), BIOME3 (Haxeltine and Prentice 1996), and IBIS (Foley et al. 1996). Input datasets for biogeography models mainly include latitude, mean monthly temperature, wind speed, solar radiation, and soil properties such as texture and depth. 40 4.3.3 Expert-based systems 45 Knowledge-based systems generally refer to expert system approaches to prediction or decision support. An expert is someone who has achieved high knowledge on a subject through her/his life experience (Kuhnert, Martin, & Griffiths 2010), therefore being a reliable source of information in a specific domain. Expert knowledge-based species-habitat Page 22 of 48 First Order Draft 5 10 15 20 IPBES Deliverable 3c DO NOT CITE relationships are used extensively to guide conservation planning, particularly when data are scarce (Iglecia et al. 2012). Expert knowledge is quite commonly utilized in conservation science (Janssen et al. 2010). Ecological expert knowledge has been frequently incorporated in aquatic habitat suitability model to link environmental conditions to the quantitative habitat suitability of aquatic species (Mouton et al. 2009). Research on species distribution modelling, which incorporated expert knowledge into modelling process is relatively limited. In a study on “finessing atlas data for species distribution models”, Niamir et al. (2011) incorporated existing knowledge into a Bayesian expert system to estimate the probability of a species being recorded at a finer than the original atlas data to predict a bird species distribution. They noted that knowledge-based species distribution maps of finer scale using a hybrid MaxEnt/expert system had a higher discriminative capacity than conventional approach even though such an approach might be limited to well-known species. In a study on “tradeoffs of different types of species occurrence data for use in systematic conservation planning”, Rondinini et al. (2006) noted that the geographic range data of species generated by expert knowledge had advantage of avoiding the potential propagation of errors simplifying data processing steps. However, eliciting expert information involves dealing with multiple expert judgements, with different sources of biases in the elicited information and with uncertainty around expert estimates (Martin et al. 2012). For instance, expertise may vary geographically, with relevant information restricted to the region of interest of the experts (Murray et al. 2009). 4.3.4 Hybrid modelling 25 30 35 40 Hybrid models combine multiple modelling approaches to represent complex, integrated systems of human and biophysical components (Parrot et al. 2011). These models are highly generally data driven and serve to aid in pattern extraction and knowledge synthesis, providing an important link between data sources and decision support systems. Developments of hybrid models take a pathway where some of the ecological processes defining the ecological system under study are (i.e. realized niche) are modelled explicitly (process-based or mechanistically), while others are kept in the correlative niche model. Hybrid approaches derive from the interest to balance realism and flexibility in model building with limited knowledge. However, this approach is not without problems. The way different models are integrated into hybrid approaches often is a difficult issue. Gallien et al. (2010) indicate that one of the current limitations of these hybrid approaches is the form and strength of the relationship between habitat suitability and demographic parameters. Changes in habitat suitability are normally integrated with population processes by limiting carrying capacity. Furthermore, how ecological processes (e.g., growth, dispersal, and thermal tolerance) respond to environmental changes is unclear, and is often assumed to be unimodal or linear. Nonlinear functional response could make the model more complex. More importantly, most functional responses used in the model are assumed, rather than confirmed from experiments. Fortunately, when simple occupancy models are used, model integration becomes easier: cells that become unsuitable force or induce extinction, whereas colonization is allowed in those cells that become suitable only. Page 23 of 48 First Order Draft 5 10 15 IPBES Deliverable 3c DO NOT CITE Broadly speaking, mechanisms determining ecosystem dynamics can be related to the ecology of species, species interactions and evolutionary processes (Lavergne et al. 2010). Any biological process of interest should have an explicit link with the components formulated in the model. However, this link does not need to be one-to-one (Lurgui et al 2014). The implementation of these processes in the model may be carried out in a wide variety of ways spanning a broad range of complexities from cellular automata (e.g. Iverson et al. 2004; Smolik et al. 2010), metapopulation models (e.g. Wilson et al. 2009), structured metapopulation models (e.g., Akçakaya 2000) to spatially-explicit population models (e.g., Cabral & Schurr 2010), individual-based models (Grimm & Railsback 2005) and reaction-diffusion models (e.g., Wikle 2003; Hui et al. 2011). Models with emergent dynamics may also include species interactions (e.g., Albert et al. 2008), or abiotic processes included via feedbacks (e.g., wildfires vs. vegetation growth; Grigulis et al. 2005). Emergent dynamics can be spatially auto correlated or not. For example, a spread model for an invasive species will generate autocorrelated spatial patterns (e.g., Merow et al. 2011), whereas a module simulating local birth and death may generate temporal dependency but may not generate auto correlated spatial structures by itself. 4.4 Modelling biodiversity feedbacks and issues across temporal and spatial scales. 20 25 30 Both human and non-living environmental drivers influence biodiversity through a number of processes. In turn, biodiversity exerts feedbacks on both systems (Fig. 4.6). Considering the feedbacks is important because they are susceptible to cause non-linearity in the interaction dynamics. This is especially important as non-linearity can potentially lead to whole-system bifurcations (e.g. regime shift) when forcing variables such as atmospheric CO2 evolve beyond observed levels. Changes in biodiversity interact with different drivers of biodiversity change (i.e. climate change, disturbance regimes such as forest fires, invasive and pests, and ecosystem processes) over different temporal and spatial scales. Changes in biodiversity and range shifts (plant traits) can influence climate at global and regional scales. For instance, General Circulation Models (GCMs) based simulations indicate that widespread replacement of deep-rooted tropical trees by shallow-rooted pasture grasses would reduce evapotranspiration and lead to a warmer, drier climate (Shukla et al. 1990). Similarly the replacement of snowcovered tundra by a dark conifer canopy at high latitudes may increase energy absorption sufficiently to act as a powerful positive feedback to regional warming (Foley et al 1994). Page 24 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE Figure 4.5 Schematic diagram of the interactions between biodiversity, the human system and the nonliving environment. The figure represents the feedbacks between biodiversity and the drivers to biodiversity change and their interactions (Source: Chapin et al 2000). 5 10 15 20 25 30 Besides, changes in biodiversity are known to induce changes in the microclimate. Such changes in microclimate may further impact the biodiversity. For example, in a boreal forest ecosystems, where soil temperatures have a strong influence on nutrient supply and productivity, the presence of moss, which reduces heat flux into the soil, provides stability to the permafrost leading to the low rates of nutrient cycling (Von Cleve et al 1991). As fire frequency increases in response to high-latitude warming, moss biomass decline, the stability of permafrost declines, nutrient supply increases, and the species compositions of forests are further altered. Chapin et al. (2010) outlines the possible interactions and feedbacks between biodiversity, biodiversity change drivers and the interactions between these drivers and suggests that anthropogenic activities (arrow 1 on Figure 4.5) cause environmental and ecological changes of global significance (2). By a variety of mechanisms, these global changes impact biodiversity, and changing biodiversity feeds to the susceptibility to species invasions (3, purple arrows). Changes in biodiversity can have direct consequences for ecosystem services and as well as human economic and social activities (4) In addition, changes in biodiversity can influence ecosystem processes (5) and ecosystem services (6). These further feedbacks to alter biodiversity (7, red arrow). Global changes may also directly affect ecosystem processes (8, blue arrows). Chapin et al. (2010) further conclude that “depending on the circumstances, the direct effects of global change may be either stronger or weaker than effects mediated by changes in diversity”. Changes in biodiversity influence annual rates of primary productivity (NPP) and decomposition of organic matters in soils. Increased NPP production relates to increased carbon sequestration from the atmosphere, and thereby provides negative feedback to Page 25 of 48 First Order Draft 5 10 IPBES Deliverable 3c DO NOT CITE climate warming. Similarly, increased decomposition of soil organic matters and litter relates to a loss of carbon from the biosphere to the atmosphere, and thereby provides a positive feedback to climate warming. Several studies using experimental species assemblages conclude “annual rates of primary productivity and nutrient retention increase with increasing plant species richness, but saturate at a rather low number of species” (Tillman et al 1996, Hector et al 1999). Salonius (1981) concludes that microbial richness can lead to increased decomposition of organic matter. Further, particular species can have strong effects on ecosystem processes by directly mediating energy and material fluxes or by altering abiotic conditions that regulate the rates of these processes. Introduction of the deep-rooted salt cedar (Tamarix sp.) to the Mojave and Sonoran Deserts of North America increased the water and soil solutes accessed by vegetation, enhanced productivity, and increased surface litter and salts. This inhibited the regeneration of many native species, leading to a general reduction in biodiversity. 15 4.4.1 Tools for assessing feedbacks 20 25 30 35 40 45 Feedbacks between drivers and biodiversity or ecosystem levels usually involve high level of complexity in the models because changes in state variables at different levels (either biological or biological and other) should be able to interact and cause emergent dynamics. Changes in biodiversity for instance can impact disturbance regimes such as fire, which in turn is strongly determined by climate (Pausas and Keely 2009) and fire-suppression efforts (Brotons et al. 2013). As an illustration, several species of nutritious but flammable grasses were introduced to the Hawaiian Islands to support cattle grazing. Some of these grasses spread into protected woodlands, where they caused a 300-fold increase in the extent of fire. Most of the woody plants, including some endangered species, are eliminated by fire, whereas grasses rebound quickly (D’Antonio et al 1992). Similar increases in the ecological role of fire resulting from grass invasions have been widely observed in the Americas, Australia, and elsewhere in Oceania. Species diversity is also known to reduce the probability of outbreaks by 'pest' species or invasives by diluting the availability of their hosts. This phenomenon particularly decreases host-specific diseases (Burdon, 1993), plant-feeding nematodes (Wasilewska, 1995) and consumption of preferred plant species (Bertness and Leonard, 1997). Further, biodiversity characteristics can impact the ability of exotic species to invade communities through either the influence of traits of resident species or some cumulative effect of species richness. Early theoretical models and observations of invasions on islands indicate that species-poor communities would be more vulnerable to invasions because they offered more empty niches (Elton, 1958). Fire disturbance models at the landscape scale have been used to integrate different processes and model feedbacks between vegetation and disturbances regimes. The LANDIS for instance model simulates forest succession, disturbance (including fire, wind, harvesting, insects), climate change, and seed dispersal across large landscapes and allows tracking the spatial distribution of discrete tree and shrub species (Scheller et al. 2007). If feedbacks are to be incorporated across scales or very different domains of applications, approaches explicitly integrating different models should be developed. One example of this complexity is exemplified by the use of Integrated Assessment Models (IAMs, see Figure 4.7). Page 26 of 48 First Order Draft 5 10 15 20 IPBES Deliverable 3c DO NOT CITE In the IPCC Third Assessment Report (IPCC, 2001) integrated assessment was defined as “an interdisciplinary process that combines, interprets, and communicates knowledge from diverse scientific disciplines from the natural and social sciences to investigate and understand causal relationships within and between complicated systems”. It is generally agreed that there are two main principles to integrated assessment, i.e. integration over a range of relevant disciplines; and the provision of information suitable for decision-making (Harremoes and Turner, 2001). IAMs thus aim to describe the complex relations between environmental, social and economic drivers that determine current and future state of the ecosystem and the effects of climate change, in order to derive policy-relevant insights (Vuuren et al., 2009). One of the essential characteristics of integrated assessment is the simultaneous consideration of the multiple dimensions of environmental problems. IAMs typically describe the cause-effect chain of climate change from economic activities and emissions to changes in climate and related impacts on e.g. ecosystems, human health and agriculture, including some of the feedbacks between these elements. In order to make their construction and use tractable, many IAMs use relatively simple equations to capture relevant phenomena. This simplification is most obvious for the climate system and carbon cycle, which, in many IAMs, consists of only a few equations (Goodess et al. 2003). However, the behaviour of these components can have a significant impact on IAM results and the quality of policy advice, with the possibility of simplifications in the earth system projections leading to imprecision (or even error) in projecting impacts and costs of mitigation. Figure 4.6 General structure of an IAM. 25 30 Over the last decade, IAMs have expanded their coverage in terms of land use and terrestrial carbon cycle representation, non-CO2 gases and air pollutants and by looking into specific impacts of climate change. Some IAMs have a stronger focus on economics, such as multisectoral computable general equilibrium models that are combined with climate modules and models focussed on cost–benefit analysis. Whereas, other IAMs are more focussed on the physical processes in both the natural system and the economy (integrated structural models/biophysical impact models). Few examples of IAMs are: IMAGE, DICE, FUND and MERGE. The basic structure of IMAGE 2.4 version is described below (see Figure 4.7). The key drivers of change, population and the macro-economy, can be derived from various external Page 27 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE and internal sources. For macro-economic drivers the exogenous source depends on the study in which the model is applied. 5 10 However, as IAMs aim to integrate different aspects of the environment, they run the risk of becoming extremely complex. Thus, the developers of such models have to make decisions about the focus of their study and how they wish to express the impacts they are attempting to estimate, whether it is through the reporting of physical changes in emissions, shifts in landuse activity or mortality rates, or through cost-benefit analysis of damages resulting from climate change (Goodess et al., 2003). The data requirements for these IAMs are also large and not always feasible. Figure 4.7 Framework of IMAGE 2.4. Source: Bouwman et al., 2006 4.5 Accounting for uncertainty 15 20 Biodiversity and ecosystems are under stress due to drivers, and the stress is likely to increase in the future due to drivers such as climate change, habitat modification, exploitation, and pollution. Policy making on biodiversity and ecosystem functioning may have to take place based on current knowledge. This must be done recognizing that uncertainty is associated with all modelling, due to limitations of data, representation of processes, and resolution of the ecosystem scale. However, policy makers have to take decisions even in the face of uncertainty, to act on the drivers in order to conserve ecosystems and biodiversity. Page 28 of 48 First Order Draft 5 10 15 20 25 30 IPBES Deliverable 3c DO NOT CITE No discussion of modelling frameworks would be complete without consideration of uncertainty (Leung et al 2012). Uncertainty in modelling of biodiversity and ecosystem processes under changing climate and land-use dynamics arises from multiple factors. For example, existing impact assessment studies including the biophysical and integrated assessment model (IAM) based studies, generally tend to examine the ‘means’ of the impacts probability distribution, neglecting the low-probability, high impact tails of the distribution (Marten et al 2013; Ackerman 2010; Weitzman 2009). Further, impact projections generally model linear climate change and inadequately account for factors such as non-linearity and tipping elements in the climate system, and irreversibility of the impacts – which are difficult to model (Whiteman et al. 2013). Impact studies generally focus on single-sector or single region-based assessments. The potential interactions, among sectors and regions, which can adversely impact the biodiversity and ecosystems, are not adequately included in the quantitative estimates (Warren 2011). Additionally, impact assessment models generally leave out the natural processes and feedbacks that are difficult to model at current state of knowledge, even though these processes may cause large impacts. For example, pest attack and fire dynamics in terrestrial and forest ecosystems are often not included in the biophysical impact assessment processes. Similarly the ambient policy and management practices and socio-economic stresses leading to degradation of natural resources are also not included in majority of the sectorial impact assessment models. Key human related issues such as armed conflicts, migrations and loss of cultural heritage has a lot of potential to impact the natural ecosystems, however impact assessment models do not include these stresses related to the human systems (Hope 2013; Theisen 2013). IAMs based economic analyses of impacts are generally conservative as these studies make optimistic assumptions about the scale and effectiveness of adaptation (Marten et al 2013; Hope 2013). Fischling et al. (2007) conclude that the most advanced tool to assess the impact of climate change on vegetation dynamics at global scale include DGVMs. 4.5.1 Sources of uncertainty Link et al. (2012) and Leung et al. (2013) highlighted six major sources of uncertainty confronting ecosystem modellers (Figure 4.8). Page 29 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE Figure 4.8 A conceptual diagram of the flow of information and actions in a typical LMR management system. Rectangles represent components of the system, solid arrows indicate flows of information and actions between components, and ellipses represent major sources of uncertainty (Link et al. 2012) 5 10 15 20 25 30 1) Natural variability Natural variability or stochasticity includes biological differences among individuals within a population, differences among populations within a community, changes in spatial distributions through time, density-dependent or independent variation in a vital rate, seasonal or interannual variability in realized environmental conditions, or shifts in productivity regimes. Natural variability increases ecosystem model uncertainty by reducing the precision of parameter estimates. 2) Observation error Observation error is inevitable when studying organisms in either a single species or an ecosystem context (e.g., Morris and Doak 2002; Ives et al. 2003 as cited in Link et al. 2012). The environmental characteristics of a particular area (even those that we can measure fairly accurately) are difficult to relate directly to the full experience of mobile organisms that move into and out of that area. Thus, natural variability can actually exacerbate observation error. Observation error adds uncertainty to ecosystem models through reduced precision, misspecified parameter distributions, and biased parameter estimates. 3) Structural complexity The structural complexity of a model arises from many factors, such as the number of parameters it includes; the number of ecosystem components and processes it simulates; the temporal scale; the nonlinearities, log effects, thresholds, and cumulative effects incorporated in those processes; and whether or not it includes features such as spatial dynamics or stochasticity (Fulton et al., 2003 as cited in Link et al. 2012). Structurally complex ecosystem models are gaining in use, in part due to improved computing capabilities and also to the intricate, multi-sector, cross-disciplinary questions commonly being addressed in ecosystemPage 30 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE based management. Thus, models have optimal levels of complexity (Walters 1986, Fulton et al., 2003). 5 10 15 20 25 30 35 40 45 Epistemic uncertainty contains parameter uncertainty, model (structural) uncertainty reflected in the model structural complexity and data/observation error. Epistemic uncertainty reflects our level of knowledge about a system, and can be reduced with additional information. We build verbal (qualitative) or mathematical (quantitative) models to represent our understanding of processes underlying a system. We use empirical data to parameterise the model. However, given stochastic elements present, we never have perfect estimates of underlying parameters and processes. Furthermore, given that models are our conceptual abstractions of the real world, even the basic structure of the model is uncertain. In addition, there will always be some uncertainty in our data sources, which may propagate through the risk model. 4) Inadequate communication Inadequate communication (or linguistic uncertainty) relates to the difficulty of effectively conveying information between scientists, managers and stakeholders. Communication problems range from the subtle to the obvious (Peterman, 2004 as cited in Link et al. 2012). When communication is ineffective, information is lost, which can manifest itself as uncertainty in many ways (Martin et al. 2012). 5) Unclear management objectives There are numerous obstacles to the development of clear objectives. Many issues related to resource use or conservation are poorly understood and are not amenable to clear objectives, or involve vague constructs that are difficult to link to performance metrics (sensu Smith et al. 1999 as cited in Link et al. 2012) 6) Outcome uncertainty Outcome uncertainty arises when realized in situ values deviate from expected values derived from a model. It is also referred to as “implementation error” or “implementation uncertainty” because it is commonly associated with differences between a management goal and the implementation of the management plan (Rice and Richards 1996; Dichmont et al. 2006; Holt and Peterman 2006; Dorner et al. 2009 as cited in Link et al. 2012). Outcome uncertainty may also be connected to other forms of uncertainty described above. For example, unclear management objectives or other forms of inadequate communication could lead stakeholders to reject management plans because the stakeholders feel their best interests are being overlooked (Watson-Wright et al. 2009 as cited in Link et al. 2012). All of the ecosystem models are diverse in terms of scope and approach, but share the general feature of a large number of parameters with complex interactions. These models are necessarily built with imperfect information. Given these inevitable uncertainties, large and complex ecosystem models must be evaluated through sensitivity analyses before their output can be effectively applied to conservation problems (Hilborn and Mangel 1997; Saltelli et al 2000b; Regan et al. 2002; Clark 2003; Harwood and Stokes 2003; Pielke and Conant 2003; Tang et al. 2006 as cited in McElhany et al. 2010). Uncertainty in climate scenarios arises from Page 31 of 48 First Order Draft 5 10 15 20 25 30 35 40 IPBES Deliverable 3c DO NOT CITE different greenhouse gas emission storylines and from differences between climate models even if driven with the same greenhouse gas emission scenario (e.g., Buisson et al. 2010 as cited in McElhany et al. 2010). This can be partly addressed by using climate change scenario data from several emission scenarios but also by using results from multi‐model studies (i.e. an ensemble of climate models). Process‐based models (PBMs) are important tools in forest and ecological science and widely used to assess the impacts of climate change on forest ecosystems (Landsberg 2003; Fontes et al. 2010 as cited in McElhany et al. 2010). Parameter uncertainty has received less attention in climate change impact studies. Climate change impact studies that do not integrate parameter uncertainty may over‐ or underestimate climate change impacts on forest ecosystems. Box. Uncertainty in DGVMs Process based, dynamic models capture the transient response of vegetation to a changing environment using explicit representation of key ecological processes such as establishment, tree growth, competition, death, nutrient cycling (Shugart 1984, Botkin 1993). With the adoption of DGVMs reliability of results has improved in relation to previous generations, however validation of some of these models is an ongoing exercise (e.g., Prentice et al. 2007). The large structural and parametric uncertainty in the DGVMs concerning the processes of recruitment, competition and tree mortality means that existing DGVMs produce a wide variety of predictions regarding the future strength and direction of the climate carbon cycle feedback (Thornton et al., 2007; Sitch et al., 2008). For example, some models predict catastrophic declines in the Amazon and boreal forests, while others predict relatively stable ecosystem composition and carbon storage, even with the same future climate drivers (as illustrated by Sitch et al., 2008). Model biases introduced by these uncertainties are not readily estimated because limited observations exist to constrain demographic processes under rapidly altering climates (Allen et al., 2009). For issues such as climate change effects on forest growth, there are large uncertainties in both the magnitude of future climate change, and the magnitude of various responses to a changed climate (Dixon and Wisniewski 1995). Many of the DGVMs do not yet simulate forest fires dynamically as well as pest attacks and land use change dynamics. Further, migration and seed dispersal constraints are not yet incorporated in DGVMs. Many of the DGVMs are known to over-simulate grasslands (Bonan et al. 2003) and at the same time there are suggestions that compared with other dynamic vegetation models, some DGVMs tends to simulate a fairly strong CO2 fertilization effect (Cramer et al. 2001; McGuire et al. 2001). Additionally uncertainty in the projections of biodiversity and ecosystem processes arises from uncertainty in climate projections as well, particularly in precipitation at downscaled regional levels (Chaturvedi et al 2012). Poor land use projections (including afforestation, reforestation and forest regeneration) due to anthropogenic influences further add to the uncertainty. However it is important that the uncertainty is adequately represented and acknowledged in the impact assessment studies. We suggest the following specific improvements for the future research to capture the full range of uncertainty in the impact assessments: Use of multiple climate models and use of multiple impact assessment models; Integrated modelling of the natural and production system Developing Plant functional Types for more vegetation types and at species level and for consideration of fauna. Page 32 of 48 First Order Draft 5 IPBES Deliverable 3c DO NOT CITE Similarly the results of complex models, such as lake ecosystem models, often suffer, however, from limitations due to various sources of error and uncertainty such as the initial conditions, input data, model structure, model parameters, validation data, etc. (Beck 1987 as cited in Gal 2012). Parameter uncertainty is a key issue when dealing with a complex model due to the large number of parameters and the uncertainty as to their true values (Helton et al. 2006). In order to increase the reliability of the model as a management tool it is important to estimate the degree of uncertainty surrounding those predicted relationships. There are diverse approaches for estimating and quantifying uncertainty of the various model components affecting model outcome (Walker et al. 2003; Helton et al. 2006). 10 15 20 25 Box. Uncertainty in species distributions models (SDMs). Modelling exercises should aim at minimising and when possible explicitly accounting for known sources of uncertainty and critical sources of uncertainty. Sources of uncertainty in biodiversity modelling used in a decision making context are not equivalent and therefore, a critical analysis should be performed in assessing which potential sources of uncertainty may impact the outcomes of the modelling approach (see Chapter 5 on decision making). Recent reviews on the role of uncertainty in biodiversity modelling approaches such as SDMs have indentified a number of uncertainty sources of different types and suggested solutions and good practices to deal with them (Rocchini et al. 2011). They identified significant sources of uncertainty in the SDMs modeling building approach and explicitly mention possible solutions that could be indentified as good practice. As an example uncertainty in input data cold be related to uncertainty in coordinate information or positional error associated to presence records that should be dealt with by improving screening of data previous to model building. This work strongly suggests that the collection of these uncertainty sources and the collection of possible solutions and good practices should be favoured in the future and greatly facilitate the development of new applications. 4.6 Options for reducing uncertainty Simeon et al. (XX) highlighted some considerations on reducing uncertainty in the structural form of models of ecosystem dynamics. 30 1. 2. 35 3. 4. 40 5. Identify the purpose of the modelling exercise in terms of management objectives, and the performance measures by which the attainment of these objectives will be assessed. Identify the key uncertainties about the system. This should occur during the process of assembling information and formulating models; it is necessary to identify and highlight uncertainties rather than to make assumptions that constrain the models to a single view of any important process. Develop models or parameterisations that represent plausible limits to each key uncertainty. Consider more than one basic model structure. Always include models, assumptions, and parameterisations that are not on the bounds, (and hence may be more plausible), and ensure that the choice of models, parameter values, and assumptions is balanced given the purpose of the modelling. Establish the full range of model behaviours by considering different combinations of models and parameters. Sensitivity analysis is useful at this stage to determine the importance of each source of uncertainty Page 33 of 48 First Order Draft 6. 5 7. 8. 10 9. 10. 15 11. 20 12. IPBES Deliverable 3c DO NOT CITE Consider the interaction between models and data. Specifically, do the models capture the full range of potential conditions, or just the conditions represented by the data? If the latter is the case, it may be advisable to consider adopting plausible future scenarios that extend outside of the range of historic data, as may occur, for example, under current predictions for climate change. Ensure that each model is logically consistent. For example, assumptions or fixed values for key parameters in an ecosystem model will also need to be made in models used to derive prior density functions for the ecosystem model’s input parameters. Weigh models by plausibility if information exists to do this. Ideally this weighting would be on the basis of posterior probability, but a more subjective weighting by prior probability might be necessary. Run each model multiple times to incorporate the effects of parameter uncertainty and natural variability. Avoid averaging model results unless the distribution of results suggested by all models is unimodal. If it is possible to weight models, present the results in terms of the risk that each management objective will not be met. If it is not possible to weight models, present the results in terms of the trade-offs associated with each management action for each alternative model. Make sure the assumptions and limitations of the approach are presented along with the results. 4.7 Ways forward in biodiversity and ecosystem modeling 25 30 35 Modelling biodiversity and ecosystem response to direct drivers of change adequately for predicting future outcomes is a challenge that can hardly be surmounted because of the inherent complexity of real-world processes (Araújo et al. 2005). In the quest towards models with enhanced predictability, biodiversity and ecosystem models will certainly become more complex and include different components. One of the costs inherent to increasing model complexity is the need for appropriate model parameters. Building complex models with unrealistic parameters derived from lack of knowledge or data is often less preferable to simple models with realistic parameters. Both model calibration and evaluation call for data, and whereas the need of obtaining realistic models for a world with many species is on the rise, little information on their spatial dynamics is currently available. An increase in the number of processes explicitly formulated in process-based models is leading to a larger number of parameters and unsurprisingly to less predictive capability and increase the uncertainty associated with the results (Walters 1986). Therefore, increased model complexity should be justified and match the scientific question at hand. In addition, increased model complexity will challenge the ability of modellers to communicate what is being formulated into the model and how this is being conceptually done. 40 Increased complexity in biodiversity model applications can be due to: (1) increased number of drivers acting in a model through input layers; or (2) increased number of processes explicitly considered in the model. In correlative niche models many biological processes are collapsed and implicitly included (Guisan & Thuiller 2005; Soberón 2007). The pathway of increasing the Page 34 of 48 First Order Draft 5 10 IPBES Deliverable 3c DO NOT CITE number of external inputs may lead to poor predictive performance of temporal extrapolations from correlative niche models (Vallecillo et al. 2009), because there is no way to specify or control the interaction between the processes implicitly included. In process-based models biological processes are included explicitly (Kearney & Porter 2009). This means that an increase of the number of external inputs can be handled appropriately if knowledge is available on the functional responses that relate all inputs with species performance. However, interactions between even a few processes can make such functions very complex (e.g., Rayfield et al. 2009). Therefore, the pathway of increasing the number of processes modelled externally may be either limited by the knowledge available or condemned to poor performance. 4.7.1 Handling increased complexity 15 Matching model complexity with biological objectives at hand is a major challenge in future development of biodiversity and ecosystem models. We describe here three general strategies that should help limiting model complexity. 4.7.1.1 Formulating critical biological processes 20 Avoiding unnecessary increases in model complexity requires a careful assessment of the biological processes that most directly affect species distributions at the spatial and temporal scales of interest for each particular study (Guisan & Thuiller 2005). Although there is no general recipe to select the relevant biological processes, processes related to species autoecology will always have a central role. Habitat selection and population dynamics in species level models may be formulated with more or less detail, but are fundamentally important to predict species distribution dynamics (Willis et al. 2009; Kunstler et al. 2011). 25 30 35 40 Moreover, movement has been recently identified as one of the most important biological processes because it allows the modelling of emergent spatial dynamics and constrains species distributions (Soberón & Nakamura 2009). However, movement is a function not only of individual species’ characteristics, but also the structure of the landscape (or seascape) over which dispersal is occurring, including the presence of natural barriers or fragmentation of habitats (Nathan & Muller-Landau 2000; Schurr et al. 2007). At the community level, biotic interactions have also been identified as important but their importance varies across species (generalist species are less dependent upon interactions with other species than specialists) and scales (interactions are more likely to be determinant at finer scales). Despite its importance, biotic interactions are likely to constitute one of the most difficult challenges faced in biodiversity models. There are some examples that have explicitly simulated the effects of competition in the distributions of individual species (e.g. Lischke et al. 2006; Albert et al. 2008), but complex sets of interactions among species are seldom considered in modelling exercises (Araújo et al. 2011; Guisan & Rahbek 2011). Finally, plasticity and evolutionary processes are rarely considered in the context of species distribution dynamics but they may result critical for species whose short generation time allows them to rapidly adapt to environmental changes (e.g., Kearney et al. 2009). Page 35 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE 4.7.1.2 Hierarchical modular modelling approaches 5 10 Biological processes should be formulated explicitly and internally (i.e. process based models) only when they are critical for the questions at hand. The remaining processes can be modelled externally and formulated into the model by means of input spatial layers or parameters modified by additional modelling frameworks (Smith et al. 2001). Such an approach may facilitate the structure of species distribution models by allowing sub-models to be plugged into one another (e.g., McRae et al. 2008). In this modular structure, the upper levels provide external contextual information (and hence external dynamics) to the lower ones. Hierarchical modular structures have the advantage of: (1) more easily integrate across different spatial and temporal scales (e.g. to downscale the results of processes formulated at higher levels) (e.g., Del Barrio et al. 2006); and (2) assess the levels of uncertainty added at each stage (Larson et al. 2004, Chisholm & Wintle 2007). However, modularity may be limited for those target species that modify their environment or interact with other biotic entities (Midgley et al. 2010). 15 4.7.1.3 Comparing models with different levels of complexity 20 Research is needed to compare the outputs of models with different degrees of complexity in light of validation data appropriate for the process or driver under study (Roura-Pascual et al. 2010). Only in this case will it be possible to build a body of reference regarding the minimum acceptable levels of complexity to analyze a given problem. 4.7.1.4 Model communication: making use of standard protocols 25 30 35 40 In the quest for more reliable predictions, SDM applications are moving beyond simple correlative models towards a wider range of process-based modeling strategies, thus converging with developments in other modeling disciplines such as population and metapopulation dynamics. This trend will likely complicate the interpretation and communication of model structure and functionality in a similar way to the problem identified more generally in other fields such as agent based modeling (Grimm et al. 2006, Grimm et al. 2010). One way to better communicate the origin of dynamics in SDMs would be the general application of standard protocols for model communication such as the ODD (Overview, Design concepts, and Detail, Grimm et al. 2006). The ODD protocol forces the modeler to be explicit in the critical issues we have identified in our conceptual review. First, the Overview section of the ODD protocol is a conceptual description of the model that includes (1) the specific purpose of the model; (2) what are basic biological entities and state variables used in model formulation; and (3) which are the processes explicitly included in the model. In line with the ODD protocol, the conceptual description of a SDM should include the biological question in hand (section 1 in Overview), a detailed description of target state variable and drivers included (section 2 in Overview) and the biological processes of interest (section 3 in Overview). The second main component in the ODD protocol tackles the design concepts defining the model formulation. In this section emergence is especially important allowing the identification of which system-level phenomena truly emerge from the explicitly formulated processes (emergent dynamics) and which are merely imposed from the inputs (imposed dynamics). When considering specific models from scientific applications, it will be important to see whether it is possible to tease apart (i) what was explicitly included in the model vs. Page 36 of 48 First Order Draft 5 10 IPBES Deliverable 3c DO NOT CITE what was imported from other sources and how; and (ii) what was actually modelled in the model vs. what was considered by the authors to be modelled, and what is the link. Finally, there is a need for approaches that combine the rigor of the small-scale studies precision with the breadth of broad-scale assessments (Naidoo and Ricketts 2006; Egoh et al. 2008; and Nelson et al. 2008 for some initial attempts). Spatially explicit values of biodiversity across landscapes that might inform land-use and management decisions are still lacking. Without quantitative assessments, and some incentives for landowners to preserve biodiversity, this tends to be ignored by land-management decisions. Similarly, rights-based fisheries have shown to have positive impacts on management of marine resources (Costello et al. 2008). Without information on the impacts of land-use management practices on biodiversity, it is impossible to design policies or payment programs that will provide the desired ecosystem services. 4.7.2 Conclusions 15 20 25 30 35 40 Modelling is the only option to policy makers to assess the implications of different drivers on biodiversity and ecosystem functioning, especially in the future context. The best examples are from climate change, where Global Circulation Models and Earth System Models help understand the role of greenhouse gas emission driver on future climate. Similarly DGVMs assist policy makers on the impact of projected climate change driver on terrestrial ecosystems and biodiversity. Modelling complex systems is characterized by inherent limitations since all processes and drivers and their linkage to biodiversity or ecosystem functions cannot be quantified and included. Further no single model that exists today covers all potential biophysical and socio-economic drivers, their inter-relationships and feedbacks. But policy-making requires robust model projections of impacts of drivers on biodiversity and ecosystem functioning. All the limitations and multiplicity of models was presented in this chapter. In the coming decades drivers such as climate change, land use and pollution will become more important. Thus scientific community has to develop strategies to address the limitations of current models and reduce uncertainty involved. Some potential examples of strategies include the following: • Formation of Model inter-comparison group, similar to CMIP - Coupled Model Intercomparison Project Phase 5 (http://cmip-pcmdi.llnl.gov/cmip5/) or ISIMIP Inter-Sectoral Impact Model Intercomparison Project (https://www.pikpotsdam.de/research/climate-impacts-and-vulnerabilities/research/rd2-crosscutting-activities/isi-mip) projects, involving a large number of modelling groups working on biodiversity and ecosystem modelling • Development of Integrated models that can be applied at sea/landscape or ecosystem level to assess the impact of drivers on biodiversity and ecosystem functioning. These integrated models should consider both bio-physical and socio-economic drivers at scales relevant to decision making. • Apply multiple models, and get a range of outputs to policy makers, rather than giving single model outputs, since there are multiple future pathways. • Need to generate data for different ecosystems at multiple locations and multiple scales by forming research networks. Page 37 of 48 First Order Draft • • 5 • IPBES Deliverable 3c DO NOT CITE Initiate long term monitoring studies to assess the response of ecosystems to changing drivers such as climate change, pollution and hydrological changes (http://unesdoc.unesco.org/images/0009/000938/093876eo.pdf). Develop protocols for modelling drivers impacting biodiversity and ecosystem functions. Encourage more collaboration between modelling groups and field ecologists. Page 38 of 48 First Order Draft IPBES Deliverable 3c DO NOT CITE References 5 10 15 20 25 30 35 40 45 Ackerman, F., Stanton, E.A., Bueno, R (2010) Fat tails, exponents, extreme uncertainty: Simulating catastrophe in DICE. Ecological Economics 69 (2010) 1657–1665 Akçakaya, H., Radeloff, V. & Mladenoff, D. (2004). Integrating landscape and metapopulation modeling approaches: viability of the sharp-tailed grouse in a dynamic landscape. Conservation Biology, 18:526-537. Alkemade R, et al. (2009) GLOBIO3: A framework to investigate options for reducing global terrestrial biodiversity loss. Ecosystems (NY) 12:374-390. Araújo, M. B., & New, M. (2007). Ensemble forecasting of species distributions. Trends in ecology & evolution, 22(1), 42-47. Araújo, M. B., & Peterson, A. T. (2012). Uses and misuses of bioclimatic envelope modeling. Ecology, 93(7), 1527-1539. Araújo, M. B., Alagador, D., Cabeza, M., Nogués‐Bravo, D., & Thuiller, W. (2011). Climate change threatens European conservation areas. Ecology letters, 14(5), 484-492. Araújo, M. B., and R.G. Pearson. (2005). Equilibrium of species' distributions with climate. Ecography 28, 693-695. Arrhenius, O. (1921). Species and area. Journal of Ecology, 9, 95–99. Ayres, R.U. 1989. Industrial metabolism and global change. Int. Soc. Sci. J. 121, 363-73. Baird D.J., Rubach M.N. & Van den Brink P.J. (2008), Trait-based ecological risk assessment (TERA): the new frontier? Integrated Environmental Assessment and Management. 4, 2–3. Bellard, C., Bertelsmeier, C., Leadley, P., Thuiller, W. & Courchamp, F. (2012). Impacts of climate change on the future of biodiversity. Ecol Lett, 15, 365–377. Berkeley, S. A., Hixon, M. A., Larson, R. J., & Love, M. S. (2004). Fisheries sustainability via protection of age structure and spatial distribution of fish populations. Fisheries, 29(8), 23-32. Bonan GB, Levis S, Sitch S, Vertenstein M, Oleson KW (2003) A dynamic global vegetation model for use with climate models: concepts and description of simulated vegetation dynamics, Global Change Biol. 9:1543–1566. Botkin DB (1993) Forest dynamics: An Ecological Model. Oxford University Press. Oxford. Bouwman AF (eds. Kram T, Klein Goldewijk K), 2006, Integrated Modelling of Global Environmental Change. An Overview of IMAGE 2.4, Report no.500110002, ISBN-9069601516. Brotons, L., Aquilué, N., De Cáceres, M., Fortin, M.J. & Fall, A. (2013). How fire history, fire suppression practices and climate change affect wildfire regimes in Mediterranean landscapes. Plos One. Bugmann H (1994) On the Ecology of Mountainous Forests in a Changing Climate: A Simulation Study Ph.D. Thesis No. 10'638. Swiss Federal Institute of Technology, Zurich (ETHZ), Switzerland 258 pp. Cardinale, B. J., Duffy, J. E., Gonzalez, A., Hooper, D. U., Perrings, C., Venail, P., et al. (2012). Biodiversity loss and its impact on humanity. Nature, 486(7401), 59-67. Cardinale, B. J., Wright, J. P., Cadotte, M. W., Carroll, I. T., Hector, A., Srivastava, D. S., et al. (2007). Impacts of plant diversity on biomass production increase through time because of species complementarity. Proceedings of the National Academy of Sciences, 104(46), 18123-18128. Cariboni J., Galtelli D., Liska R., Saltelli A., (2007). The role of sensitivity analysis in ecological modelling. Ecological Modelling. 203, 167-182. Chapin, F. S., Zavaleta, E. S., Eviner, V. T., Naylor, R. L., Vitousek, P. M., Reynolds, H. L., et al. (2000). Consequences of changing biodiversity. Nature, 405 (6783), 234–242. Chaturvedi RK, Jaideep Joshi, Mathangi Jayaraman, G. Bala, N.H. Ravindranath (2012). Multi-model climate change projections for India under Representative Concentration Pathways. Current Science, 103(7):791-802 Cheung, W. W., Sarmiento, J. L., Dunne, J., Frölicher, T. L., Lam, V. W., Palomares, M. D., ... & Pauly, D. (2013). Shrinking of fishes exacerbates impacts of global ocean changes on marine ecosystems. Nature Climate Change, 3(3), 254-258. Page 39 of 48 First Order Draft 5 10 15 20 25 30 35 40 45 IPBES Deliverable 3c DO NOT CITE Cheung, W.W.L., Lam, V.W., Pauly, D., 2008. Dynamic bioclimate envelope model to predict climateinduced changes in distribution of marine fishes and invertebrates. In: Cheung, W.W.L., Lam, V.W.Y., Pauly, D. (Eds.), Modeling Present and Climate-shifted Distribution of Marine Fishes and Invertebrates, vol. 163. Fisheries Centre, University of British Columbia: Fisheries Centre Research Reports, pp. 5–45. Cheung, W.W.L., Lam, V.W.Y., Sarmiento, J.L., Kearney, K., Watson, R. & Pauly, D. (2009). Projecting global marine biodiversity impacts under climate change scenarios. Fish and Fisheries, 10, 235– 251. Christensen, V., & Pauly, D. (1992). ECOPATH II --- a software for balancing steady-state ecosystem models and calculating network characteristics. Ecological Modelling, 61(3-4), 169–185. doi:10.1016/0304-3800(92)90016-8 Christensen, V., & Walters, C. J. (2004). Ecopath with Ecosim: methods, capabilities and limitations. Ecological Modelling, 172(2-4), 109–139. Chun J. H. and C. B. Lee. (2013). Assessing the Effects of Climate Change on the Geographic Distribution of Pinus densiflora in Korea using Ecological Niche Model. Korean Journal of Agricultural and Forest Meteorology, 15(4), 219-233. Clark WC (2010) Sustainable development and sustainability science. In: Levin, S. A., Clark, W. C. (Eds), Toward a science of sustainability: Report from Toward a Science of Sustainability Conference. Princeton: Princeton University Press. Collie, J. S., Botsford, L. W., Hastings, A., Kaplan, I. C., Largier, J. L., Livingston, P. A., et al. (2014). Ecosystem models for fisheries management: finding the sweet spot. Fish and Fisheries. Concepción, E. D., & Díaz, M. (2011). Field, landscape and regional effects of farmland management on specialist open-land birds: Does body size matter?. Agriculture, Ecosystems & Environment, 142(3), 303-310. Costello, C., Gaines, S.D., Lynham, J., 2008. Can Catch Shares Prevent Fisheries Collapse? Science 321, 1678–1681. doi:10.1126/science.1159478 Cramer et al (2001) Response of terrestrial ecosystem structure and function to CO2 and climate change: results from six dynamic global vegetation models, Global Change Biology 7:357-373. Dale, V. H., & Beyeler, S. C. (2001). Challenges in the development and use of ecological indicators. Ecological indicators, 1(1), 3-10. Davis, A. J., Jenkinson, L. S., Lawton, J. H., Shorrocks, B., & Wood, S. (1998). Making mistakes when predicting shifts in species range in response to global warming. Nature, 391(6669), 783-786. de Chazal, J. & Rounsevell, M.D.A. (2009). Land-use and climate change within assessments of biodiversity change: A review. Global Environmental Change, 19, 306–315. Dengler, J. (2009). Which function describes the species–area relationship best? A review and empirical evaluation. Journal of Biogeography, 36(4), 728-744. Desrochers, R. E., Kerr, J. T., & Currie, D. J. (2011). How, and how much, natural cover loss increases species richness. Global Ecology and Biogeography, 20(6), 857-867. Detlef P. van Vuuren, Jason Lowe, Elke Stehfest, Laila Gohar, Andries F. Hof, Chris Hope, Rachel Warren, Malte Meinshausen, Gian-Kasper Plattner, 2009, How well do integrated assessment models simulate climate change?, Springer, Climate change, DOI 10.1007/s10584-009-9764-2 Dixon RK and Wisniewski J (1995) Global forest systems: An uncertain response to atmospheric pollutants and global climate change?, Water, Air, Soil Pollution 85:1:101–110. Dormann, C.F., Schymanski, S.J., Cabral, J., Chuine, I., Graham, C., Hartig, F., et al. (2012). Correlation and process in species distribution models: bridging a dichotomy. J. Biogeogr., 39, 2119–2131. Drakare, S., Lennon, J. J., & Hillebrand, H. (2006). The imprint of the geographical, evolutionary and ecological context on species–area relationships. Ecology letters, 9(2), 215-227. Egoh B, Reyers B, Rouget M, et al. 2008. Mapping ecosystem service for planning and management. Agri Ecosyst Environ 127: 135–40. Page 40 of 48 First Order Draft 5 10 15 20 25 30 35 40 45 IPBES Deliverable 3c DO NOT CITE Eldridge, D. J., & Freudenberger, D. (2005). Ecosystem wicks: woodland trees enhance water infiltration in a fragmented agricultural landscape in eastern Australia. Austral Ecology, 30(3), 336-347. Elith, J., & Leathwick, J. R. (2009). Species distribution models: ecological explanation and prediction across space and time. Annual Review of Ecology, Evolution, and Systematics, 40(1), 677. Elith, J., Kearney, M., & Phillips, S. (2010). The art of modelling range-shifting species. Methods in Ecology and Evolution, 1, 330-342. Elith, J., Phillips, S. J., Hastie, T., Dudík, M., Chee, Y. E., & Yates, C. J. (2011). A statistical explanation of MaxEnt for ecologists. Diversity and Distributions, 17(1), 43-57. Ewers, R.M., Kapos, V., Coomes, D.A., Lafortezza, R., Didham, R.K., 2009. Mapping community change in modified landscapes. Biological Conservation 142, 2800-2872. Faith DP, Ferrier S, Williams KJ. 2008. Getting biodiversity intactness indices right: ensuring that ‘biodiversity’ reflects‘diversity’. Glob Chang Biol 14:207-21. Ferrier, S., Manion, G., Elith, J., Richardson, K., 2007. Using generalized dissimilarity modelling to analyse and predict patterns of beta diversity in regional biodiversity assessment. Diversity and Distributions 13, 252-264. Ferrier, S., Watson, G., Pearce, J., Drielsma, M., 2002. Extended statistical approaches to modelling spatial pattern in biodiversity in northeast New South Wales. I.Species-level modelling. Biodiversity and Conservation 11, 2275-2307. Ferrier,S., Guisan,A. 2006. Spatial modelling of biodiversity at the community level. Journal of Applied Ecology 43, 393-404. Ferrier,S.(2002) Mapping spatial pattern in biodiversity for regional conservation planning: where to from here? Systematic Biology, 51, 331-363. Fischlin A, Midgley GF, Price JT, Leemans R, Gopal B, Turley C, Rounsevell MDA, Dube OP, Tarazona J and Velichko AA (2007) Ecosystems, their properties, goods, and services. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change.(eds.Parry ML, Canziani OF, Palutikof JP, van der Linden PJ and Hanson CE, Cambridge University Press, Cambridge, 211272. Foley AJ, Levis S, Costa MH, Cramer W and Pollard D (2000) Incorporating dynamic vegetation cover within global climate models ecological applications, Ecological Society of America 10(6):16201632. Fordham, D. A., Akçakaya, H. R., Brook, B. W., Rodríguez, A., Alves, P. C., Civantos, E. et al. (2013). Adapted conservation measures are required to save the Iberian lynx in a changing climate. Nature Climate Change, 3(10), 899-903 Fulton, E.A., Fuller, M., Smith, A.D.M., Punt, A.E. 2005. Ecological indicators of the ecosystem effects of fishing: final report. Australian Fisheries Management Authority Report, R99/1546, 239 pp. Gaikwad, J., Wilson, P. D., & Ranganathan, S. (2011). Ecological niche modeling of customary medicinal plant species used by Australian Aborigines to identify species-rich and culturally valuable areas for conservation. Ecological Modelling, 222(18), 3437-3443. Gal, G., V. Makler-Pick, and N. Shachara (2012). Dealing with uncertainty in ecosystem model scenarios: application of single model ensemble approach. International Environmental Modelling and Software Society (iEMSs) 2012 International Congress on Environmental Modelling and Software. http://www.iemss.org/society/index.php/iemss 2012 proceedings Gallien, L., Münkemüller, T., Albert, C.H., Boulangeat, I. & Thuiller, W. (2010). Predicting potential distributions of invasive species: where to go from here? Diversity and Distributions, 16, 331– 342. Garcia, S. M., Kolding, J., Rice, J., Rochet, M. J., Zhou, S., Arimoto, T., ... & Smith, A. D. M. (2012). Reconsidering the consequences of selective fisheries. Science, 335(6072), 1045-1047 Page 41 of 48 First Order Draft 5 10 15 20 25 30 35 40 45 IPBES Deliverable 3c DO NOT CITE Gaston, K.J., Davies, R.G., Orme, C.D.L., Olson, V.A., Thomas, G.H., Tzung-Su, D.,Rasmussen, P.C., Lennon, J.J., Bennett, P.M., Owens, I.P.F., Blackburn, T.M., 2007.Spatial turnover in the global avifauna. Proceedings of the Royal Society of London Series B 274, 1567-1574. Goodess CM, Hanson C, Hulme M, Osborn TJ (2003) Representing climate and extreme weather events in integrated assessment models: a review of existing methods and options for development. Integr Assess 4:145–171 Graedel, T.E. and Crutzen, P.J. 1990. Atmospheric trace constituents. In: The Earth as Transformedbv HumanAction. Turner, B.L., II, Clark, W.C., Kates, R.W., Richards, J.F., Mathews, J.T. and Meyer, W.B. (eds). Cambridge University Press, Cambridge, p. 295-311. Guilhaumon, F., Gimenez, O., Gaston, K. J., & Mouillot, D. (2008). Taxonomic and regional uncertainty in species–area relationships and the identification of richness hotspots. Proceedings of the National Academy of Sciences, 105, 15458–15463. Guisan, A., Tingley, R., Baumgartner, J.B., Naujokaitis-Lewis, I., Sutcliffe, P.R., Tulloch, A.I.T., et al. (2013). Predicting species distributions for conservation decisions. Ecol Lett, 16, 1424–1435. Guisan, A.,Zimmermann,N.E. (2000). Predictive habitat distribution models in ecology. EcologicalModelling. 135, 147-186. Gustafson, E. J. (2013). When relationships estimated in the past cannot be used to predict the future: using mechanistic models to predict landscape ecological dynamics in a changing world. Landscape ecology, 28(8), 1429-1437. Halpern, B. S., Walbridge, S., Selkoe, K. A., Kappel, C. V., Micheli, F., D'Agrosa, C., et al. (2008). A global map of human impact on marine ecosystems. Science, 319(5865), 948–952. Handa, I. T., Aerts, R., Berendse, F., Berg, M. P., Bruder, A., Butenschoen, O., et al. (2014). Consequences of biodiversity loss for litter decomposition across biomes. Nature, 509(7499), 218-221. Hanski, I., Zurita, G. A., Bellocq, M. I., & Rybicki, J. (2013). Species–fragmented area relationship. Proceedings of the National Academy of Sciences, 110(31), 12715-12720. Harfoot, M. B., Newbold, T., Tittensor, D. P., Emmott, S., Hutton, J., Lyutsarev, V., ... & Purves, D. W. (2014). Emergent Global Patterns of Ecosystem Structure and Function from a Mechanistic General Ecosystem Model. PLoS biology, 12(4), e1001841. Harremoes P, Turner RK (2001) Methods for integrated assessment. Reg Environ Change 2:57–65 Haxeltine A and Prentice IC (1996) BIOME3, an equilibrium terrestrial biosphere model based on ecophysiological constraints, resource availability and competition among plant functional types, Global Biogeochemical Cycles 10.4:693-710. Hector et al. 1999. Helton, J.C., Johnson, J.D., Sallaberry, C., Storlie, C.B. (2006). Survey of sampling-based methods for uncertainty and sensitivity analysis. Reliability Engineering & System Safety, 91, 1175-1209. Hickler, T., Vohland, K., Feehan, J., Miller, P. A., Smith, B., Costa, L., ... & Sykes, M. T. (2012). Projecting the future distribution of European potential natural vegetation zones with a generalized, tree species‐based dynamic vegetation model. Global Ecology and Biogeography, 21(1), 50-63. Hilbert,D.W.& Ostendorf,B(2001) The utility of artificial neural networks for modelling the distribution of vegetation in past, present and future climates. Ecological Modelling,146, 311-327. Hope C (2013) Critical issues for the calculation of the social cost of CO2: why the estimates from PAGE09 are higher than those from PAGE2002. Climatic Change. 117:531-543 Iglecia, M. N., J. A. Collazo, and A. J. McKerrow. 2012. Use of occupancy models to valuate expert and knowledge-based species-habitat relationships. Avian Conservation and Ecology. 7(2), 5. http://dx.doi.org/10.5751/ACE-00551-070205. Iverson, L. and Anantha M. Prasad (2001). Potential Changes in Tree Species Richness and Forest Community Types following Climate Change. Ecosystems, 4, 186-199. Jacobs, S., Burkhard, B., Van Daele, T., Staes, J. & Schneiders, A. (2015). Ecological Modelling. Ecol Model, 295, 21–30. Page 42 of 48 First Order Draft 5 10 15 20 25 30 35 40 45 IPBES Deliverable 3c DO NOT CITE Jager, T., & Klok, C. (2010). Extrapolating toxic effects on individuals to the population level: the role of dynamic energy budgets. Philosophical Transactions of the Royal Society B: Biological Sciences, 365(1557), 3531-3540. Janssen, J.a.E.B., Krol, M.S., Schielen, R.M.J., Hoekstra, A.Y. & De Kok, J.L. (2010) Assessment of uncertainties in expert knowledge, illustrated in fuzzy rule-based models. Ecological Modelling, 221, 1245-1251. Jennings, S., Mélin, F., Blanchard, J.L., Forster, R.M., Dulvy, N.K., Wilson, R.W., 2008. Global-scale predictions of community and ecosystem properties from simple ecological theory. Proceedings of the Royal Society of London B 275, 1375–1383. Jennings, S., Pinnegar, J.K., Polunin, N.V.C., Warr, K.J., 2002. Linking size-based and trophic analyses of benthic community structure. Mar Ecol Prog Ser 226, 77–85. Jørgensen, S. E., 1994. Fundamentals of Ecological Modelling. Developments in Environmental Modelling, 2nd edition, vol. 19. Elsevier, Amsterdam Kaplan JO et al. (2003) Climate change and Arctic Ecosystems: Modeling, Paleodata-Model comparisons, and future projections, Journal of Geophysical Research-Atmosphere 108:8171. Neilson RP (1995) A model for predicting continental-scale vegetation distribution and water balance, Ecological Applications 5:362-385. Kates, R.W., Tumer, B.L., II and Clark, W.C. 1990. The great transformation. In: The Earth as Transformed by Human Action. Turner, B.L., II, Clark, W.C., Kates, R.W., Richards, J.F., Mathews, J.T. and Meyer, W.B. (eds). Cambridge University Press, Cambridge, p. 1-17. Kearney, M., Simpson, S.J., Raubenheimer, D. & Helmuth, B. (2010a). Modelling the ecological niche from functional traits. Philosophical Transactions of the Royal Society B: Biological Sciences, 365, 3469–3483. Kearney, M.R., Wintle, B.A. & Porter, W.P. (2010b). Correlative and mechanistic models of species distribution provide congruent forecasts under climate change. Conservation Letters, 3, 203–213. Koleff, P., Gaston, K.J. & Lennon, J.K. (2003) Measuring beta diversity for presence-absence data. Journal of Animal Ecology, 72, 367-382. Kuhnert, P.M., Martin, T.G. & Griffiths, S.P. (2010) A guide to eliciting and using expert knowledge in Bayesian ecological models. Ecology Letters, 13, 900–914. Lavergne, S., Mouquet, N., Thuiller, W. & Ronce, O. (2010). Biodiversity and Climate Change: Integrating Evolutionary and Ecological Responses of Species and Communities. Annual Review of Ecology, Evolution, and Systematics, 41, 321–350. Legendre, P., Lapointe, F.-J. & Casgrain, P. (1994) Modeling brain evolution from behavior: a permutational regression approach. Evolution, 48, 1487-1499. Lenhardt, P. P., Schäfer, R. B., Theissinger, K., & Brühl, C. A. (2013). An expert-based landscape permeability model for assessing the impact of agricultural management on amphibian migration. Basic and Applied Ecology, 14(5), 442-451 Lennon, J.J., Koleff, P., Greenwood, J.J.D., Gaston, K.J., 2001. The geographical structure of British bird distributions: diversity, spatial turnover and scale. Journal of Animal Ecology 70, 966-979. Leung, B., Roura-Pascual, N., Bacher, S., Heikkilä, J., Brotons, L., Burgman, M.A., et al. (2012). TEASIng apart alien species risk assessments: a framework for best practices. Ecol Lett, 15, 1475–1493. Link, J. S., Ihde, T. F., Harvey, C. J., Gaichas, S. K., Field, J. C., Brodziak, J. K. T., et al. (2012). Dealing with uncertainty in ecosystem models: The paradox of use for living marine resource management. Progress in Oceanography, 102, 102–114. doi:10.1016/j.pocean.2012.03.008 Loh J, Green RE, Ricketts T, Lamoreux J, Jenkins M, Kapos V,Randers J. 2005. The Living Plant Index: using species population time series to track trends in biodiversity. Phil Trans R Soc B 360:289-295. Loreau M., Mouquet N., Gonzalez A. (2003) Biodiversity as spatial insurance in heterogeneous landscapes. PNAS 100(22): 12765–12770. Page 43 of 48 First Order Draft 5 10 15 20 25 30 35 40 45 IPBES Deliverable 3c DO NOT CITE Luoto, M.,Poyry,J.,Heikkinen,R.K.,Saarinen,K., (2005). Uncertainty of bioclimate envelope models based on the geographical distribution of species. GlobalEcology and Biogeography 14, 575–584. Lurgi, M., Brook, B.W., Saltré, F. & Fordham, D.A. (2014). Modelling range dynamics under global change: which framework and why? Methods in Ecology and Evolution, in press. Mace, G. M., Norris, K., & Fitter, A. H. (2012). Biodiversity and ecosystem services: a multilayered relationship. Trends in ecology & evolution, 27(1), 19-26 Majer JD, Beeston G. 1996. The Biodiversity Integrity Index: an illustration using ants in Western Australia. Conserv Biol 10:65-73. Malcolm, J. R., Liu, C., Neilson, R. P., Hansen, L., & Hannah, L. E. E. (2006). Global warming and extinctions of endemic species from biodiversity hotspots. Conservation biology, 20(2), 538-548. Marten, A.L., Robert E. Kopp., Kate C Shouse., Charles W. Griffiths., Elke L. Hodson., Elizabeth Kopits., Bryan K. Mignone& Chris Moore & Steve C. Newbold & Stephanie Waldhoff & Ann Wolverton (2013) Improving the assessment and valuation of climate change impacts for policy and regulatory analysis. Climatic Change. 117:433–438 Martin, T.G., Burgman, M. a., Fidler, F., Kuhnert, P.M., Low-Choy, S., McBride, M.F. & Mengersen, K. (2012) Eliciting expert knowledge in conservation science. Conservation biology, 26, 29–38. Maury, O. & J.C. Poggiale (2013) From individuals to populations to communities: A dynamic energy budget model of marine ecosystem size-spectrum including life history diversity, Journal of Theoretical Biology, 324, 52-71 McElhany, P., E.A. Steel, D. Jensen, K. Avery, N. Yoder, C. Busack and B. Thompson. (2010). Dealing with uncertainty in ecosystem models: lessons from a complex salmon model. Ecological Applications. 20, 465-482. McGuire AD, Melillo JM, Joyce LA, Kicklighter DW, Grace AL, Moore III B and Vorosmarty CJ (1992) Interactions between carbon and nitrogen dynamics in estimating net primary productivity for potential vegetation in North America, Global Biogeochemical Cycles 6:101-124. McGuire AD, Sitch S, Clein JS, Dargaville R, Esser G et al (2001) carbon balance of the terrestrial biosphere in the twentieth century: analyses of CO2, climate and land-use effects with four process-based ecosystem models, Global Biogeochemical Cycles 15:183-206. McInerny, G.J. & Etienne, R.S. (2012). Ditch the niche - is the niche a useful concept in ecology or species distribution modelling? J. Biogeogr., 39, 2096–2102. McMahon, S.M., Harrison, S.P., Armbruster, W.S., Bartlein, P.J., Beale, C.M., Edwards, M.E., et al. (2011). Improving assessment and modelling of climate change impacts on global terrestrial biodiversity. TREE, 26, 249–259. Merow, C., Smith, M. J., & Silander, J. A. (2013). A practical guide to MaxEnt for modeling species’ distributions: what it does, and why inputs and settings matter. Ecography, 36(10), 1058-1069. Midgley, G., Davies, I., Albert, C., Altwegg, R. & et al. (2009). BioMove an integrated platform simulating the dynamic response of species to environmental change. Ecography, 33:1-5. Molnár, P. K., Derocher, A. E., Klanjscek, T., & Lewis, M. A. (2011). Predicting climate change impacts on polar bear litter size. Nature Communications, 2, 186. Morin, X., & Lechowicz, M. J. (2008). Contemporary perspectives on the niche that can improve models of species range shifts under climate change. Biology Letters, 4(5), 573-576. Morin, X., & Thuiller, W. (2009). Comparing niche-and process-based models to reduce prediction uncertainty in species range shifts under climate change. Ecology, 90(5), 1301-1313. Mouton, A.M., De Baets, B., Goethals, P.L.M. (2009). Knowledge-based versus data-driven fuzzy habitat suitability models for river management. Environmental Modelling & Software, 24, 982-993. Muñoz, M.E.S. Giovanni, R., Siqueira, M.F., Sutton, T., Brewer, P., Pereira, R.S., Canhos, D.A.L. & Canhos, V.P. (2011). "openModeller: a generic approach to species' potential distribution modelling". GeoInformatica. 15, 111–135. Page 44 of 48 First Order Draft 5 10 15 20 25 30 35 40 45 IPBES Deliverable 3c DO NOT CITE Murphy, J.M, D M. H. Sexton, D N. Barnett, G S. Jones, MJ. Webb, M Collins & D A. Stainforth, 2004, Quantification of modeling uncertainties in a large ensemble of climate change simulations, Nature 430, 768-772 (12 August 2004) Murray, J. V., Goldizen, A.W., O’Leary, R.A., McAlpine, C. a., Possingham, H.P., Choy, S.L. & O’Leary, R. a. (2009) How useful is expert opinion for predicting the distribution of a species within and beyond the region of expertise? A case study using brush-tailed rock-wallabies Petrogale penicillata. Journal of Applied Ecology, 46, 842–851. Naidoo R and Ricketts TH. 2006. Mapping the economic costs and benefits of conservation. PLoS Biol 4: 2153–64 Niamir, A., Skidmore, A.K., Toxopeus, A.G., Muñoz, A.R.M. & Real, R. (2011) Finessing atlas data for species distribution models. Diversity and Distributions. 17, 1173–1185. Noss, R. F. (1990). Indicators for monitoring biodiversity: a hierarchical approach. Conservation biology, 4(4), 355-364. Parrott, L. (2011). Hybrid modelling of complex ecological systems for decision support: Recent successes and future perspectives. Ecological Informatics, 6, 44–49. Parton WJ, Scurlock JMO, Ojima DS, Gilmanov TG, Scholes RJ, Schimel DS, Kirchner T, Menaut J-C, Seastedt T, Moya EG, Kamnalrut A, Kinyamario JL (1993) Observations and modeling of biomass and soil organic matter dynamics for the grassland biome worldwide, Global Biogeochemical Cycles 7:785-809. Pausas, J.G. & Keeley, J.E. (2009). A Burning Story: The Role of Fire in the History of Life. BioScience, 59, 593–601. Payne, K. and Stockwell, D. (2006). GARP Modelling System User's Guide and Technical Reference, http://landshape.org/enm/garp-modelling-system-users-guide-and-technical-reference/ Pearson, R.G. and T.P. Dawson. (2003). Predicting the impacts of climate change on the distribution of species: Are bioclimate envelope models useful? Global Ecology and Biogeography, 12, 361-371. Pearson, R.G. C.J. Raxworthy, M. Nakamura, and A.T. Peterson. (2007). Predicting species' distributions from small numbers of occurrence records: A test case using cryptic geckos in Madagascar. Journal of Biogeography 34, 102-117. Pearson, R.G. T.P. Dawson, P.M. Berry, and P.A. Harrison. (2002). Species: A spatial evaluation of climate impact on the envelope of species. Ecological Modelling 154, 289-300. Peppler-Lisbach, C. & Schröder, B. (2004) Predicting the species composition of Nardus stricta communities by logistic regression modelling. Journal of Vegetation Science, 15, 623-634. Pereira, H. M., & Daily, G. C. (2006). Modeling biodiversity dynamics in countryside landscapes. Ecology, 87, 1877–1885 Pereira, H.M., Leadley, P.W., Proenca, V., Alkemade, R., Scharlemann, J.P.W., Fernandez-Manjarres, J.F., et al. (2010). Scenarios for Global Biodiversity in the 21st Century. Science, 330, 1496–1501. Pereira, H.M., Navarro, L.M. & Martins, I.S. (2012). Global Biodiversity Change: The Bad, the Good, and the Unknown. Annu. Rev. Environ. Resourc., 37, 25–50. Peterson G., Allen C.R., Holling C.S. (1998) Ecological resilience, biodiversity, and scale. Ecosystems 1: 6– 18 Peterson, A. T., & Vieglais, D. A. (2001). Predicting Species Invasions Using Ecological Niche Modeling: New Approaches from Bioinformatics Attack a Pressing Problem. BioScience, 51(5), 363-371. Peterson, A. T., Soberón, J., Pearson, R. G., Anderson, R. P., Martínez-Meyer, E., Nakamura, M., Araújo, M.B. (2011). Ecological niches and geographic distributions (MPB-49). Princeton University Press. Peterson, A.T. (2006). Uses and requirements of ecological niche models and related distributional models. Biodiversity Informatics, 3. 59-72. Peterson, A.T. M.A. Ortega-Huerta, J. Bartley, V. Sanchez-Cordero, J. Soberon, R.W. Buddemeier, and D.R.B. Stockwell. (2002). Future projections for mexican faunas under global climate change scenarios. Nature 416, 626-629. Page 45 of 48 First Order Draft 5 10 15 20 25 30 35 40 45 IPBES Deliverable 3c DO NOT CITE Phillips, S. J., & Dudík, M. (2008). Modeling of species distributions with Maxent: new extensions and a comprehensive evaluation. Ecography, 31(2), 161-175. Polovina, J. J. (1984). Model of a coral reef ecosystem. Coral Reefs, 3(1), 1–11. Prentice IC, Bondeau A, Cramer W, Harrison SP, Hickler T, Lucht W, Sitch S, Smith B and Sykes MT (2007) Dynamic Global Vegetation Modelling: Quantifying Terrestrial Ecosystem Responses to Largescale Environmental Change (eds. J.G. Canadell, D.E. Pataki and L.F. Pitelka) Springer-Verlag, Berlin, Terrestrial Ecosystems in a ChangingWorld, 175-192. Proença, V., & Pereira, H. M. (2013). Species–area models to assess biodiversity change in multi-habitat landscapes: The importance of species habitat affinity. Basic and Applied Ecology, 14(2), 102-114. Rahbek, C., Gotelli, N. J., Colwell, R. K., Entsminger, G. L., Rangel, T. F. L., & Graves, G. R. (2007). Predicting continental-scale patterns of bird species richness with spatially explicit models. Proceedings of the Royal Society B: Biological Sciences, 274(1607), 165-174 Rajiv K Chaturvedi, 2012, Climate change mitigation and adaptation in Indian forests, Lambert Academic Publishers, Germany ISBN: 978-3-659-13500-2. Rajiv K Chaturvedi, Ranjith Gopalakrishnan, Mathangi Jayaraman, Govindasamy Bala, NV Joshi, Raman Sukumar, NH Ravindranath, 2011, Impact of climate change on Indian forests: a dynamic vegetation modeling approach, Mitigation and adaptation strategies for global change, Vol 6, issue 2 Raxworthy, C. J., Martinez-Meyer, E., Horning, N., Nussbaum, R. A., Schneider, G. E., Ortega-Huerta, M. A., & Peterson, A. T. (2003). Predicting distributions of known and unknown reptile species in Madagascar. Nature, 426(6968), 837-841. Rocchini, D., Hortal, J., Lengyel, S., Lobo, J.M., Jimenez-Valverde, A., Ricotta, C., et al. (2011). Accounting for uncertainty when mapping species distributions: The need for maps of ignorance. Prog Phys Geogr, 35, 211–226. Rockström J, Steffen W, Noone K, Person A, Chapin FS, Lambin EF, et al. (2009) A safe operating space for humanity. Nature 461: 472-475. Rodrigues M, Oliveira A, Costa M, Fortunato AB, Y. Zhang (2009) Sensitivity analysis of an ecological model applied to the Ria de Aveiro. J Coastal Research. SI 56, 448-452. Rondinini C. Kerrie, A. Wilson, Luigi Boitani, Hedley Grantham and Hugh P. Possingham (2006). Tradeoffs of different types of species occurrence data for use in systematic conservation planning. Ecology Letters. 9, 1136-1145. Rosenzweig, M. L. (1995). Species diversity in space and time. Cambridge: Cambridge University Press. Running SW, Hunt Jr. ER (1993) Generalization of a Forest Ecosystem Process Model for Other Biomes, BIOME-BGC, and an Application for Global-scale Models, (eds. Ehleringer JR, Field C). Academic Press, San Diego, CaliforniaScaling processes between leaf and landscape levels.141-158. Salonius 1981 Sarmiento, J. L. S., and N. Gruber (2006), Ocean Biogeochemical Dynamics, 503 pp., Princton Univ. Press, Princeton, N. J. Scheller, R. M., J. B. Domingo, B. R. Sturtevant, J. S. Williams, A. Rudy, E. J. Gustafson, and D. J. Mladenoff. 2007. Design, development, and application of LANDIS-II, a spatial landscape simulation model with flexible spatial and temporal resolution. Ecological Modelling 201 (3-4): 409-419. Scholes RJ, Biggs R. 2005. A biodiversity intactness index. Nature 434:45-9. Shin, Y.J., Cury, P.M., 2004. Using an individual-based model of fish assemblages to study the response of size spectra to changes in fishing. Canadian Journal of Fisheries and Aquatic Sciences 61, 414– 431. Shugart HH (1984). A theory of forest dynamics: The Ecological Implication of Forest Succession Models. Springer-Verlag, New York. Page 46 of 48 First Order Draft 5 10 15 20 25 30 35 40 45 IPBES Deliverable 3c DO NOT CITE Shugart HH, Urban DL (1989). Factors Affecting the Relative Abundances of Tree Species. Eds.Grubb PJ, Whittaker JB, Towards a more exact ecology. 30th Symposium of the British Ecological Society, Blackwell, Oxford -, 249-273. Sitch S (2000) The Role of Vegetation Dynamics in the Control of Atmospheric CO2 Content, PhD Thesis, Lund University. Allen CD (2009) Climate-induced forest dieback: an escalating global phenomenon?, Unasylva 60:231-232. Smith, J.U., & Smith, P. (2007). Environmental modelling. An introduction. Oxford University Press, Oxford Snowling, S., and Kramer, J. (2001). Evaluating modelling uncertainty for model selection. Ecological Modelling. 138, 17-30. Soetaert, K., & Herman, P. M. (Eds.). (2009). A practical guide to ecological modelling: using R as a simulation platform. Springer. Steffen W, Crutzen PJ, McNeill JR (2007) The anthropocene: Are humans now overwhelming the great forces of nature? Ambio 36(8): 614-621. Steffen W, Sanderson A, Tyson PD, Jäger J, Matson P, Moore III B, et al. (2004) Global change and the earth system: A planet under pressure. The IGBP global change series. Berlin, Germany: SpringerVerlag. Steffen WL, Walker BH, Ingram JSI and Koch GW (1992) Global change and terrestrial ecosystems: The operational plan. International Geosphere-Biosphere Programme. Stock, C.A. et al. (2011). On the use of IPCC-class models to assess the impact of climate on living marine resources. Progress in Oceanography, 88, 1–27. Stockwell, D.R.B. and I.R. Noble. (1992). Induction of sets of rules from animal distribution data: a robust and informative method of data analysis. Mathematics and Computers in Simulation, 33, 385390. Theisen, OM., Nils Petter Gleditsch, Halvard Buhaug (2013) Is climate change a driver of armed conflict? Climatic Change. 117:613-625 Thornton, P. E., & Zimmermann N. E. (2007). An Improved Canopy Integration Scheme for a Land Surface Model with Prognostic Canopy Structure. Journal of Climate. 20(15), 3902 - 3923. Thuiller, W., Münkemüller, T., Lavergne, S., Mouillot, D., Mouquet, N., Schiffers, K., et al. (2013). A road map for integrating eco-evolutionary processes into biodiversity models. Ecol Lett, 16, 94–105. Tikkanen, O. P., Punttila, P., & Heikkilä, R. (2009). Species–area relationships of red‐listed species in old boreal forests: a large‐scale data analysis. Diversity and Distributions, 15(5), 852-862. Tillman et al. 1996 Tittensor, D. P., Micheli, F., Nyström, M., & Worm, B. (2007). Human impacts on the species–area relationship in reef fish assemblages. Ecology Letters, 10(9), 760-772. Triantis, K. A., Guilhaumon, F., & Whittaker, R. J. (2012). The island species–area relationship: biology and statistics. Journal of Biogeography, 39(2), 215-231. UNEP. (2004) Decisions adopted by the conference of the parties to the convention on biological diversity at its seventh meeting (UNEP/CBD/COP/7/21/Part 2), decision VII/30 (CBD2004). http://www.biodiv.org/decisions/. Vallecillo S, Brotons L, Thuiller W. (2009) Dangers of predicting bird species distributions in response to land-cover changes: the role of dynamic processes. Ecological Applications 19(2): 538-549. van Vuuren, D. P., Sala, O. E., & Pereira, H. M. (2006). The future of vascular plant diversity under four global scenarios. Ecology and Society, 11, 25. Walker, W.E., Harremoes, P., Rotmans, J., Van der Sluijs, J., Van Asselt, M.,Janssen, P., Von Krauss, M.P.K. (2003). Defining uncertainty: a conceptual basis for uncertainty management in modelbased decision support. Integrated Assessment, 4, 5-17 Page 47 of 48 First Order Draft 5 10 15 20 IPBES Deliverable 3c DO NOT CITE Walters, C., 1986. Adaptive management of renewable resources. McMillan, New York [Reprint edition available from Blackburn Press, 2001]. Walters, C., Pauly, D., Christensen, V., & Kitchell, J. F. (2000). Representing density dependent consequences of life history strategies in aquatic ecosystems: EcoSim II. Ecosystems, 3(1), 70–83. Walters, C., Christensen, V., Walters, W., & Rose, K. (2010). Representation of multi-stanza life histories in Ecospace models for spatial organization of ecosystem trophic interaction patterns. Bulletin of Marine Science, 86, 439–459. Warren R (2011) The role of interactions in a world implementing adaptation and mitigation solutions to climate change. Phil Trans Roy Soc A 369:217–241 Watling, L. (2005). The global destruction of bottom habitats by mobile fishing gear. In E. A. Norse & L. B. Crowder, Marine conservation biology the science of maintaining the seas biodiversity (pp. 198– 210). Washington, DC: 496 pp. Weitzman M. 2009. On Modeling and Interpreting the Economics of Catastrophic Climate Change. Review of Economics and Statistics, 91(1):1-19 Whiteman G, Chris Hope & Peter Wadhams. 2013. Climate science: Vast costs of Arctic change. Nature 499, 401–403 Wiersma, Y.F., Urban, D.L., 2005. Beta diversity and nature reserve system design in the Yukon, Canada. Conservation Biology 19, 1262-1272. Wohlgemuth, T. (1998) Modelling floristic species richness on a regional scale: a case study in Switzerland. Biodiversity and Conservation, 7, 159-177. Page 48 of 48