Maintaining Habitat Connectivity for Conservation

by

Bronwyn Rayfield

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Ecology and Evolutionary Biology University of Toronto

Copyright c 2009 by Bronwyn Rayfield Abstract

Maintaining Habitat Connectivity for Conservation

Bronwyn Rayfield

Doctor of Philosophy

Department of Ecology and Evolutionary Biology

University of Toronto

2009

Conserving biodiversity in human-dominated landscapes requires protecting networks of ecological reserves and managing the intervening matrix to maintain the potential for species to move among them. This dissertation provides original insights towards

(1) identifying areas for protection in reserves that are critical to maintain biodiversity and (2) assessing the potential for species’ movements among habitat patches in a re- serve network. I develop and test methods that will facilitate conservation planning to promote viable, resilient populations through time.

The first part of this dissertation tests and develops reserve selection strategies that protect either a single focal species in a dynamic landscape or multiple interact- ing species in a static landscape. Using a simulation model of boreal forest dynamics,

I test the effectiveness of static and dynamic reserves to maintain spatial habitat re- quirements of a focal species, American marten (Martes americana). Dynamic reserves

improved upon static reserves but re-locating reserves was constrained by fragmenta-

tion of the matrix. Management of the spatial and temporal distribution of land-uses

in the matrix will therefore be essential to retain options for re-locating reserves in the

future. Additionally, to include essential consumer-resource interactions into reserve

selection, a new algorithm is presented for American marten and its two primary prey

species. The inclusion of their interaction had the benefit of producing spatially aggre-

gated reserves based on functional species requirements.

ii The second part of this dissertation evaluates and synthesizes the network-theoretic approach to quantify connectivity among habitat patches or reserves embedded within spatially heterogeneous landscapes. I conduct a sensitivity analysis of network-theoretic connectivity analyses that derive least-cost movement behavior from the underlying cost surface which describes the relative ecological costs of dispersing through different landcover types. Landscape structure is shown to affect how sensitive least-cost graph connectivity assessments are to the quality (relative cost values) of landcover types. I develop a conceptual framework to classify network connectivity statistics based on the component of habitat connectivity that they quantify and the level within the network to which they can be applied. Together, the combination of reserve design and network connectivity analyses provide complementary insights to inform spatial planning deci- sions for conservation.

iii Acknowledgements

My research on the protection of habitat networks would not have been possible with- out a very supportive social network of collaborators, friends, and family. I want to thank, first and foremost, my advisor and mentor Dr. Marie-Jos´eeFortin. Working with her has taught me so much about ecology, research, leadership, and life. I hope to be as good an advisor to my own students one day. I am also thankful to my com- mittee members, Dr. Don Jackson and Dr. Lisa Manne, for their thoughtful sugges- tions and ongoing encouragement. A special thank-you is reserved for my Aunt Jan, an inspirational wildlife biologist and naturalist, who showed me my first caribou and shared her passion for Canadian wilderness with me.

This research benefited greatly from my collaborations with Dr. Andrew Fall, Dr.

Atte Moilanen, and Dr. Dean Urban. Andrew and I shared many inspiring discussions about my research, particularly at the formative stages, and his sharp thinking helped to clarify and refine many ideas. Atte was a role model for efficiency in writing and he generously shared his insights into conservation planning and his conservation software

- Zonation. Dean kindly hosted me at Duke University to share with me his enthusi- asm and ideas about the many conservation applications of graph and network models.

Generous financial support for my work was provided by the Natural Sciences and

Engineering Research Council of Canada through a Postgraduate Scholarship (PGS-

A) and a Canada Graduate Scholarship (CGS) awarded to myself and a Discorvery

Grant awarded to my advisor, Dr. Fortin. Additional funding was provided by The

Department of Ecology and Evolutionary Biology at the University of Toronto.

The Department of Ecology and Evolutionary Biology at the University of Toronto has been a wonderful place to do graduate work. The whole Information Technology

Group, and Stephen Smith in particular, were invaluable. I am enormously grateful to all members of Spatial Ecology Lab (LE Lab), especially Patrick James, Aleksan- dra Polakowska, Stephanie Melles, Pilar Hernandez, Alistair MacKenzie, Josie Hughes,

iv Jonathan Ruppert, and Allan Brand. It has been a pleasure to share the daily comings and goings of grad life along with major successes and challenges with these people.

Many thanks to my lifelong friends - Emily, Melissa, Aleks, Ryan, and Colin - who never let me take myself too seriously. Lastly, thank you to my loved ones: Mom, Dad,

Sarah, Marc, Monkey, and Jamie. No words can express what you each mean to me.

v Chapter Acknowledgements

This thesis is comprised of five co-authored manuscripts that are either published, in press, submitted, or in preparation for peer-reviewed journals (Chapters 1-5). Permis- sions to use published materials in this dissertation have been obtained from the pub- lishers. Experimental design, analyses, and manuscript preparation were all carried out by the principal author and PhD candidate. Co-authors of the chapters contributed conceptual discussions, expertise in computer programming, and editing of written ma- terials. Chapter 1 introduces and provides the motivation for the dissertation. Chapter

6 concludes the dissertation and provides a starting point for future research.

1. Rayfield, B., Fortin, M.-J., Urban, D. (in prep.) A network approach for conser-

vation planning in dynamic landscape mosaics. (Chapter 1)

2. Rayfield, B., James, P., Fall, A., and Fortin, M.-J. (2008) Comparing static ver-

sus dynamic protected areas in dynamic boreal ecosystems. Biological Conserva-

tion 141:438-449. Reprinted with permission from Elsevier. (Chapter 2)

3. Rayfield, B., Moilanen, A., and Fortin, M.-J. (2009) Incorporating consumer-

resource spatial interactions in reserve design. Ecological Modelling. 220, 725-

733. Reprinted with permission from Elsevier. (Chapter 3)

4. Rayfield, B., Fortin, M.-J., Fall, A. (accepted with minor revisions March 2009;

LAND-08-1820 ) The sensitivity of least-cost habitat graphs to relative cost sur-

face values. Landscape Ecology. (Chapter 4)

5. Rayfield, B., Fortin, M.-J., Fall, A. (in prep.) Connectivity for conservation: A

framework to classify habitat network connectivity statistics. (Chapter 5)

vi Contents

1 Network approach for conservation planning 1

1.1 Introduction ...... 1

1.1.1 Static and equilibrated landscapes: Island Biogeography Theory . 2

1.1.2 Network models and dynamic landscape mosaics ...... 4

1.2 Network-theoretical insights into conservation planning and reserve design 5

1.2.1 Assessing the resilience of reserve networks ...... 5

1.2.2 Network robustness and network structure ...... 8

1.2.3 Connectivity and resilience in reserve networks ...... 10

1.3 Conclusion ...... 11

1.4 Dissertation overview ...... 12

1.4.1 Chapter 2: Dynamic reserves in a dynamic boreal forest . . . . . 12

1.4.2 Chapter 3: Consumer-resource interactions in reserves ...... 13

1.4.3 Chapter 4: Sensitivity of habitat network connectivity assessments 13

1.4.4 Chapter 5: Quantifying connectivity of habitat networks . . . . . 14

1.4.5 Chapter 6: Conclusions and future directions ...... 14

2 Comparing static versus dynamic reserves 15

2.1 Abstract ...... 15

2.2 Introduction ...... 16

2.3 Methods ...... 20

vii 2.3.1 Study area ...... 20

2.3.2 Boreal forest landscape dynamics model ...... 21

2.3.3 Protected areas ...... 25

2.4 Results ...... 32

2.5 Discussion ...... 35

3 Consumer-resource interactions in reserves 41

3.1 Abstract ...... 41

3.2 Introduction ...... 42

3.3 Methods ...... 44

3.3.1 Summary of the Zonation reserve-selection algorithm ...... 44

3.3.2 Introducing novel spatial consumer-resource interactions into Zona-

tion ...... 49

3.3.3 Case study: predator-prey interaction in the boreal forest of Qu´ebec

(Canada) ...... 51

3.3.4 Reserve-selection scenario comparison with different combinations

of species-specific and interaction connectivity layers ...... 52

3.4 Results ...... 53

3.5 Discussion ...... 56

3.6 Conclusion ...... 62

4 Sensitivity of least-cost habitat graphs 64

4.1 Abstract ...... 64

4.2 Introduction ...... 65

4.3 Methods ...... 71

4.3.1 Generation of artificial landscape spatial patterns ...... 71

4.3.2 Cost values to quantify resistance to movement ...... 74

viii 4.3.3 Graph-theoretic representations of habitat connectivity using least-

cost links ...... 75

4.3.4 Measuring the sensitivity of graphs with least-cost links ...... 79

4.4 Results ...... 79

4.5 Discussion ...... 84

4.6 Conclusion ...... 88

5 Classification of network-connectivity statistics 90

5.1 Abstract ...... 90

5.2 Introduction ...... 91

5.3 Background ...... 93

5.4 Development of methods to construct habitat networks ...... 97

5.5 Quantifying connectivity in habitat networks ...... 101

5.6 Classification framework of habitat network connectivity statistics . . . . 102

5.6.1 Network levels of analysis ...... 103

5.6.2 Components of habitat connectivity ...... 105

5.7 Missing habitat-network connectivity statistics ...... 110

5.8 Selecting network connectivity statistics ...... 113

5.9 Spatio-temporal connectivity in dynamic habitat networks ...... 113

5.10 Conclusions ...... 115

6 Conclusions 128

6.1 Thesis summary ...... 128

6.2 Future research directions ...... 133

Bibliography 136

ix List of Tables

2.1 Description of alternative protected area (PA) scenarios ...... 26

3.1 Focal species’ habitat, home range, and dispersal parameter estimates . . 46

3.2 Differences among reserve-selection scenarios ...... 54

4.1 Chronological and alphabetical presentation of connectivity studies using

a cost surface to identify least-cost routes ...... 67

4.2 Parameters used to generate spatial patterns of landscapes and their fac-

torial combinations ...... 74

4.3 Sets of relative cost values used in factorial experiment ...... 76

4.4 Effects of matrix composition (HM COV), habitat fragmentation (H FRAG),

matrix fragmentation (HM FRAG), and the relative cost values (COST)

on the mean spatial deviation of least-cost links ...... 81

5.1 Summary of the scale-specific classification framework for network statis-

tics that quantify habitat connectivity ...... 104

5.2 Topological connectivity statistics that have been applied to quantify con-

nectivity in habitat networks ...... 117

5.3 Node- and link-weighted connectivity statistics that have been applied to

quantify connectivity in habitat networks ...... 121

x List of Figures

1.1 Schematic representation of fragmented landscapes under a) the Island

Biogeography Theory (IBT) and b) the dynamic landscape mosaic model 6

2.1 Location of the Vermillion study area in the province of Qu´ebec, Canada 22

2.2 Illustration of the two steps of the PA selection process ...... 30

2.3 Mean properties of all home ranges in the study area ...... 34

2.4 Mean properties of home ranges in the static, static-core, and dynamic PAs 36

3.1 Example of the two types of interaction layers used as input for the Zona-

tion reserve-selection algorithm ...... 48

3.2 Reserves selected under five different reserve-selection scenarios ...... 57

3.3 Three cases illustrating varying degrees of co-occurrence between consumer

and resource ...... 59

4.1 Examples of the simulated landscapes at a resolution of 100 × 100 cells . 73

4.2 An illustration of a) least-cost links, b) paths, and c) habitat graphs. A

link is a route that directly connects two habitat patches (nodes) . . . . . 77

4.3 Interaction plots for the percentage of hospitable matrix (HM COV) versus

the set of relative cost values (C1 - C8) when habitat fragmentation is

either a) low (H FRAG = 0.05) or b) high (H FRAG = 0.5) ...... 82

xi 4.4 Interaction plots for the fragmentation of habitat (H FRAG) versus the

set of relative cost values (C1 - C8) plotted separately at each percentage

of hospitable matrix (HM COV; a - e) ...... 83

5.1 Number of published studies that use graph, network, or circuit theory to

quantify habitat connectivity ...... 94

5.2 Examples of different levels of analysis within a habitat network . . . . . 106

xii Chapter 1

A network approach for conservation planning in dynamic landscape mosaics

1.1 Introduction

One of the goals of conservation planning should be to maintain or restore resilient landscapes that have the capacity to absorb disturbances (Bengtsson et al., 2003), thereby maintaining the full variety of biodiversity, at all levels of organization, into the long-term future (Margules and Pressey, 2000). Achieving this goal in multi-use, dynamic landscapes (e.g., managed boreal forests; Bergeron et al., 1999) necessitates landscape-level conservation strategies that create networks of protected areas or re- serves while managing the landscape matrix surrounding those reserves (Lindenmayer and Franklin, 2002; Lindenmayer et al., 2006; Rayfield et al., 2008; Brudvig et al.,

2009). Designing effective conservation strategies requires knowledge of: (1) the nat- ural dynamics of landscapes and communities (Bengtsson et al., 2003; Leroux et al.,

2007a,b; Rayfield et al., 2008); (2) the response of individual species, interacting species,

1 Chapter 1. Network approach for conservation planning 2 and communities to landscape modification (Tewksbury et al., 2002; Fahrig, 2007; Ray-

field et al., 2009); and, (3) taxa-specific use of habitat fragments and the surround- ing matrix (McIntyre and Wiens, 1999; Bossenbroek et al., 2005; Guilhaumon et al.,

2008; Rayfield et al., accepted). If conservation efforts are successful, then ecological resilience should emerge as a quantifiable property of these landscapes over time, mea- sured as the capacity of the landscape to absorb disturbances and reorganize without transitioning to an alternate stable state (Holling, 1973; Gunderson, 2000).

The recent attention given to landscape dynamics (Cumming et al., 1996; Meir et al., 2004; Wimberly, 2006; Leroux et al., 2007a,b; Rayfield et al., 2008) and ecolog- ical resilience (Bengtsson et al., 2003; Folke, 2003; Lindenmayer et al., 2006) in con- servation theory reflects an important paradigm shift. There has been a move away from a static view of ecosystems that were thought to be in equilibrium with local re- sources and environmental conditions, towards a dynamic view of ecosystems that are continually being perturbed and that may exist in alternate stable states contingent on historical factors (Pulliam and Johnson, 2002; Beisner et al., 2003). Concurrently, the definition of resilience has progressed from a measure of the time it would take a perturbed system to return to equilibrium towards a measure of the magnitude of dis- turbance that a system can absorb before changing its stable state (i.e., its function, structure, identity, and feedbacks; Folke et al., 2004; Walker et al., 2004). The implica- tions of this paradigm shift for conservation methods are presented below.

1.1.1 Static and equilibrated landscapes: Island Biogeography

Theory

The equilibrium view of ecosystems led to one of the central theories in conservation

- the theory of island biogeography (IBT; MacArthur and Wilson, 1967; Pulliam and

Johnson, 2002). A key aspect of the IBT that holds relevance for conservation is that the local species richness on islands is held in equilibrium not only by the resources Chapter 1. Network approach for conservation planning 3 present but also by a balance between immigration and local extinctions. The empha- sis that IBT placed on immigration and emigration as a predictor of local species rich- ness was novel at the time of its publication (Pulliam and Johnson, 2002) and has since precipitated research into the causes and consequences of movement among stationary islands (and later discrete, static patches of habitat). Maintaining movement among spatially subdivided populations and resource patches has since been shown to promote the persistence of species by stabilizing demographic stochasticity (Hanski and Gag- giotti, 2004, but see Levins, 1969 for an earlier example), allowing for patch coloniza- tion (Taylor et al., 1993; Hanski, 1999a; Hanski and Ovaskainen, 2000), and sustaining allelic diversity and evolutionary potential (Barton, 1992; Couvet, 2002 but see Wright,

1943 for an earlier example).

The IBT presented a very simplistic model of movement through a uniformly in- hospitable “sea” such that islands were considered more or less isolated based on their size and their straight-line distance from a source population (such as the mainland;

Fig. 1.1a). In terrestrial landscapes, IBT has had a large impact on conservation where reserves are “islands” in a landscape matrix (Fig. 1.1a; binary treatment of habitat).

This simplistic model of movement has proven to have limited applicability in terres- trial systems where animals (even birds; B´elisle,2000) rarely move along straight-line paths in a landscape (Lindenmayer and Franklin, 2002). Indeed, the vast majority of terrestrial landscapes consist of a mosaic of land-cover types and habitat patches that provide the physical and biological conditions required for the survival, growth, and reproductive success for a particular species (Hall et al., 1997; Yapp, 1922). Habitat patches are more or less effectively isolated than straight-line distance would suggest depending on the intervening land-cover types within the matrix and the movement capability of the species (D’Eon et al., 2002). Chapter 1. Network approach for conservation planning 4

1.1.2 Network models and dynamic landscape mosaics

As most landscapes are dynamic, new paradigms were proposed to move beyond a bi- nary and static description of landscapes to a heterogeneous landscape model com- prised of dynamic habitat patches embedded within a mosaic of other dynamic land- cover types of varying qualities (Fig. 1.1b; Wimberly, 2006; Wiens, 2008). The abil- ity of a species to move among habitat patches in these dynamic landscape mosaics can be described as potential or functional habitat connectivity (Calabrese and Fagan,

2004; Fagan and Calabrese, 2006) rather than simply focusing on a structural assess- ment based on the physical configuration of habitats (i.e., habitat isolation as defined in IBT). Potential habitat connectivity in static landscape mosaics is measured as a function of the spatial structure of the landscape and estimates of the movement be- haviour of the species in response to that spatial structure (Taylor et al., 1993; With et al., 1997; Tischendorf and Fahrig, 2000b; Calabrese and Fagan, 2004; Fahrig, 2007).

In dynamic landscape mosaics, habitat connectivity will depend on both spatial and temporal habitat structure, species’ movement speed or rate of spread (Matlack and

Monde, 2004; Wimberly, 2006), and how species movement is affected by landscape composition and configuration (Wiens, 1997; Fahrig, 2007).

A network modelling framework (also called a graph-theoretic framework) is well- suited to capture our contemporary understanding of dynamic landscape mosaics (Ur- ban and Keitt, 2001; Fall et al., 2007; Brooks et al., 2008; Minor and Urban, 2008; Ur- ban et al., 2009). A network is a set of nodes that are connected to varying degrees by links that join pairs of nodes (Brandes and Erlebach, 2005). In conservation ap- plications, nodes represent either habitat patches or reserves and links represent the dispersal likelihood or frequency between nodes (Fig. 1.1b; Fall et al., 2007; Minor and

Urban, 2008). Weights are assigned to nodes and links to include additional informa- tion about their level of connectivity, such as the size of nodes or the length of links in a reserve network. Incorporating the spatial properties of nodes and links is partic- Chapter 1. Network approach for conservation planning 5 ularly important in conservation and landscape applications; nodes can be defined as two-dimensional patches with fixed spatial locations and links can be defined as geo- referenced links between nodes connecting patch perimeter to patch perimeter (O’Brien et al., 2006; Theobald, 2006; Fall et al., 2007; Rayfield et al., accepted) and traversing along least-cost routes based on the suitability of the intervening matrix (Halpin and

Bunn, 2000; O’Brien et al., 2006; Theobald, 2006; Fall et al., 2007).

1.2 Network-theoretical insights into conservation

planning and reserve design

Network theory explicitly focuses on networks that are either naturally-occurring or have arisen out of decentralized planning by independent decisions over long time peri- ods (e.g., habitat and reserve networks; Newman et al., 2006). The network-theoretic approach recognizes that networks are constantly changing over time and analyzes their dynamic structural properties (e.g., resilience; Watts, 1999; Newman et al., 2006).

Ultimately, the goal of network analyses is to understand the relationship between the structural properties of networks, their function, and their dynamical behaviour (e.g., functioning ecological communities; May, 2006; Newman et al., 2006). Hence, network theory can provide valuable insights for conservation planning in dynamic landscape mosaics in which maintaining connectivity among protected habitat patches in a re- serve network is vital to the persistence of biodiversity.

1.2.1 Assessing the resilience of reserve networks

The numerous ways in which network theory can inform different aspects of conser- vation planning and reserve design have been extolled by many authors (Urban and

Keitt, 2001; Brooks, 2003; Chetkiewicz et al., 2006; Fall et al., 2007; Minor and Ur- ban, 2008; Urban et al., 2009). However, one aspect of reserve design which has been Chapter 1. Network approach for conservation planning 6

Figure 1.1: Schematic representation of fragmented landscapes under a) the Island

Biogeography Theory (IBT) and b) the dynamic landscape mosaic model. In a) black nodes represent protected areas and the surrounding white area represents a uniformly inhospitable matrix. Solid lines represent the potential for an to move be- tween nodes. In IBT nodes are equally well connected if they are of equal size and separated by equal distance. Note the landscape and connections among patches are static through time. In b) black and white nodes represent intact and disturbed pro- tected areas respectively. Patches in grey scale represent different land-cover types in the landscape with increasing resistance for darker colours. Dashed line represents a linear barrier to movement such as a road. Solid lines represent least-cost links between nodes. Note the quality and configuration of land-cover types change through time and least-cost links between nodes reflect those changes. Chapter 1. Network approach for conservation planning 7 under-emphasized in conservation applications of network theory and which is partic- ularly important in dynamic landscape mosaics is reserve network resilience (but see

Urban and Keitt, 2001; Minor and Urban, 2008; Urban et al., 2009). Similar to ecolog- ical resilience (Walker et al., 2004), network resilience can be defined as the capacity of a network to absorb disturbance and retain essentially the same function, structure, identity, and feedbacks. When assessing network resilience, one is not concerned about the level of connection between any two particular nodes but rather that the network as a whole continues to function despite some endemic level of node or link destruction

(average behaviour of the network; Newman et al., 2006). The implication for reserve design is that if connection patterns among reserves are chosen appropriately, the re- serve network can be highly resilient to the loss of nodes and links (although it may be susceptible to certain forms of node loss such as disturbances which specifically target highly connected nodes; Albert et al., 2000).

Network resilience is often measured in terms of network robustness - the number of nodes or links that can be removed without altering the connectivity of the network

(Albert et al., 2000; Callaway et al., 2000; Jeong et al., 2001; Newman, 2003). This as- sumes that the ability of a reserve network to sustain its biodiversity relies at least in part on the maintenance of connectivity among individual reserves (Peterson, 2002).

Network robustness can be characterized in a number of ways depending on: (1) the way connectivity is measured; (2) whether nodes or links are removed; and (3) the or- der in which nodes or links are removed. Hence, reserve network resilience could be assessed in terms of how connectivity of the network responds to different types of dis- turbances (i.e., random vs. targeted) and different aspects of the disturbance regime

(i.e., spatial pattern of disturbances; Bengtsson et al., 2003).

There are many ways to measure the connectivity of a network (see Chapter 5), hence the choice of a connectivity measure should be tied to an explicit hypothesis about how the structure of the reserve network affects key ecological processes that Chapter 1. Network approach for conservation planning 8 will ultimately influence the persistence of biodiversity. For example, the characteristic path length (i.e., the mean shortest path length between any two nodes in the network) and the diameter (i.e., the maximum shortest path length between any two nodes in the network) are two related connectivity measures that have been used in robust- ness assessments of the World-Wide Web and habitat networks respectively (Albert et al., 2000; Urban and Keitt, 2001). A short characteristic path length relative to the number of nodes provides a good indication that a network has a few highly connected nodes that make it possible to access all information on the web with just a few clicks of the mouse (Watts and Strogatz, 1998; Albert and Barab´asi,1999). A short diame- ter in a habitat network relative to the number of nodes implies that the network can be traversed rapidly so that a dispersing individual can more easily access all habitat patches (Minor and Urban, 2008).

1.2.2 Network robustness and network structure

Network robustness depends on a variety of structural properties of the network (Al- bert et al., 2000; Jeong et al., 2001; Strogatz, 2001). To assess the effects of network structure on robustness, node and link deletions can either be performed on specific real-world networks (e.g., the World-Wide Web) or on abstract networks that are stochas- tically generated to produce certain structural properties (e.g., scale-free networks).

The distribution of links among nodes, or node-degree distribution, is a structural property of networks that affects the robustness of a network to the loss of random and highly connected nodes. Networks in which all nodes have a similar number of links

(i.e., Poisson node-degree distributions) will quickly become disconnected into small clusters when nodes are deleted in either a random or targeted fashion (Strogatz, 2001; but see Holland and Hastings (2008), for an example in which randomizing network structure is positively related to the dynamics and persistence of modelled predator- prey systems). However, networks in which nodes have a highly heterogeneous node- Chapter 1. Network approach for conservation planning 9 degree distribution with a tail decaying as a power-law (i.e., power-law node-degree dis- tribution also referred to as a scale-free network) are relatively robust to the deletion of random nodes, but are very sensitive to the targeted deletion of highly connected nodes or “hubs” (Albert and Barab´asi,1999; Jeong et al., 2001). Based on these gen- eral results, Minor and Urban (2008) proposed that an ideal reserve network would be a scale-free network comprised of a few large “hub” reserves connected to many smaller reserves.

Ferrari and Lookingbill (2009) have called into question the degree to which hubs that represent habitat patches or reserves in a landscape would function as true hubs from the classical network perspective described above. Classical hubs generally have much higher node degrees than the mean node degree of the network (orders of mag- nitude higher) and therefore they significantly decrease the average path length of the network and account for a disproportionately large portion of the overall connectivity of the network (Barab´asi,2002). Ferrari and Lookingbill (2009) point out that habi- tat networks typically have fewer than 103 nodes compared to networks such as the

World-Wide Web which have more than 106 nodes; hence, the classical definition of a hub may not apply to even the most connected nodes in reserve networks. Another important distinction to point out is that spatial constraints may eliminate the possi- bility of movement between very distant nodes in a reserve network whereas hyperlinks connecting webpages in the World-Wide Web have no constraints and hence links are possible between all pairs of nodes. The node degree heterogeneity in reserve networks is therefore limited due to fewer nodes and the presence of spatial constraints. Conse- quently, it may not be possible for habitat patch hubs to have node degrees that are orders of magnitude larger than the mean. Much more research is needed to character- ize the structural properties of existing and proposed reserve networks. In particular it is important to examine the consequences of the limited node-degree heterogeneity of reserve networks on their robustness. Chapter 1. Network approach for conservation planning 10

1.2.3 Connectivity and resilience in reserve networks

Network robustness is one approach to study the resilience of networks which focuses on the structural consequences of disturbances that remove nodes and links. Another way to think about the resilience of a reserve network is to consider structural prop- erties of the network that can help to balance the connectivity for desirable processes

(e.g., daily foraging movements among patches, inter-population movements to main- tain genetic mixing and maintain source-sink dynamics, seasonal migrations, and range shifts associated with global change) and undesirable processes (Schafer, 2001; Janssen et al., 2006, e.g., spread of invasive species, spread of forest fires, coherence of other types of environmental stochasticity). This approach recognizes that the consequences of the structural properties of networks depend on the function of desirable and unde- sirable flows within the network (Janssen et al., 2006).

Conservation planners have struggled for some time with this trade-off between locating reserves close enough that they allow for occasional exchange of individuals among them and far enough that they reduce the potential that a single disturbance or disease will impact them all (Schafer, 2001). Network theory can contribute to this quandary by identifying structural network properties that promote connectivity for certain processes at certain scales while breaking connectivity for others. For exam- ple, a structural property of networks that may enhance resilience is high compartmen- talization whereby nodes with many links are connected to nodes with few links and vice versa (Meli´anand Bascompte, 2002; Minor and Urban, 2008; Urban et al., 2009).

Compartmentalization creates connected sub-network groupings that are internally well connected but which are only loosely connected to other sub-network groupings.

Resilience is enhanced through compartmentalization because disturbances are con-

fined within this network structure (Meli´anand Bascompte, 2002). Minor and Urban

(2008) specified that in their idealized reserve network highly connected nodes (“hubs”) should be spatially separated to create compartmentalization that would isolate distur- Chapter 1. Network approach for conservation planning 11 bances but permit species movements across the landscape. The property of compart- mentalization should apply to reserve networks because it requires node-degree hetero- geneity but does not require that the node-degree distribution follows a power law.

1.3 Conclusion

Conservation planning in dynamic landscape mosaics can benefit from a network-based approach to modeling the consequences of natural and anthropogenic disturbances on the resilience of ecological communities persisting in reserve networks. A network per- spective focuses on the structural properties of landscapes and reserve networks that may enhance resilience such as node-degree distribution and compartmentalization.

The resilience of networks is an active area of research in many disciplines and there are many other structural properties that should be explored in the context of reserve networks such as connectance (Dunne et al., 2002), centrality (Jeong et al., 2001), and redundancy (Urban et al., 2009). Analysis of structural robustness of reserve networks should consider alternative sequences of node and link deletion that more closely mimic realistic patterns of disturbances (e.g., Srinivasan et al., 2007). Functional considera- tions regarding the spread of the disturbance and the dispersal abilities of species of in- terest can be simultaneously accounted for in reserve network structure. Examining the response of processes spreading through a given network structure simultaneously will improve our ability to estimate functional consequences for biodiversity conservation

(Janssen et al., 2006). Land-uses in the matrix surrounding reserves should be incorpo- rated into conservation planning whenever possible (Lindenmayer and Franklin, 2002;

Bengtsson et al., 2003; Brudvig et al., 2009). The matrix will be very important for maintaining dispersal routes among reserves (Tewksbury et al., 2002; Damschen et al.,

2008; Walker and Craighead, 1997; Adriaensen et al., 2003; Pinto and Keitt, 2009) and for maintaining areas that are capable of becoming viable habitat in the future (Ray- Chapter 1. Network approach for conservation planning 12

field et al., 2008). Resilient ecological communities will only emerge if the dynamics and contingencies of landscapes are incorporated into reserve network design and man- agement of the matrix, thereby maintaining connectivity among reserves for species of conservation concern while limiting connectivity for disturbances.

1.4 Dissertation overview

Designing reserve networks and quantifying connectivity among individual reserves are two research areas which are essential to inform the design of effective conservation management actions. My dissertation contributes to both of these areas of research.

I examine how the structure and dynamics of landscapes influence reserves and habi- tat connectivity. Chapters 2 and 3 focus on reserve design strategies for dynamic and static landscapes respectively where planning is based on the habitat requirements of a single indicator species (Chapter 2) or interacting species (Chapter 3). Chapters 4 and

5 focus on quantifying habitat connectivity using network theory. I first evaluate the sensitivity of the habitat network model to the definition of matrix quality (Chapter

4) and then develop a conceptual framework to classify network connectivity statis- tics based on their ecological interpretation (Chapter 5). Taken together, these four chapters provide a complementary suite of novel methods and approaches to meet bio- diversity persistence goals in reserves and improve assessments of potential connectivity among reserves.

1.4.1 Chapter 2: Dynamic reserves in a dynamic boreal forest

Reserve selection based on static habitat distributions may not ensure habitat persis- tence and species survival in dynamic landscapes. I compare the effectiveness of static reserves which are fixed in space through time with that of dynamic reserves which are periodically re-located in response to dynamic habitat. Reserves are selected based Chapter 1. Network approach for conservation planning 13 on the spatial habitat requirements of American marten (Martes americana) which is considered an indicator species for mature boreal forest biodiversity. I use a spatially- explicit landscape-disturbance model for a boreal forest region in Qu´ebec (Canada) to show that dynamic reserves do not dramatically improve upon static reserves because logging and fire disturbances in the matrix constrain options for re-locating reserves.

1.4.2 Chapter 3: Consumer-resource interactions in reserves

Despite their ecological importance, species interactions have not been explicitly in- cluded in terrestrial reserve-selection methods. I present a novel method for selecting reserves based on maintaining the connectivity between the distributions of a consumer and its resources. I illustrate the method with a conservation planning case-study of a predator, American Marten, and its two primary prey species in the same boreal for- est in Qu´ebec. Including consumer-resource interactions in reserve-selection resulted in more spatially aggregated reserves when compared to reserves selected based on species spatial habitat requirements with no interaction.

1.4.3 Chapter 4: Sensitivity of habitat network connectivity

assessments

Least-cost routes through landscape cost surfaces are commonly used to model species- specific landscape traversability in a spatially explicit manner. There is considerable uncertainty involved in estimating the cost values for each land-cover type within a landscape for a particular species. I investigate how sensitive connectivity assessments based on least-cost habitat networks are to relative differences in the cost values. I per- formed this sensitivity analysis in artificial landscapes that I generated to control the area and degree of fragmentation of each land-cover. I found that the least-cost habi- tat networks were sensitive to differences in relative cost values assigned to land-cover Chapter 1. Network approach for conservation planning 14 types and that the spatial structure of the landscape was an important determinant of the degree of network sensitivity.

1.4.4 Chapter 5: Quantifying connectivity of habitat networks

The number of connectivity statistics for habitat networks can easily overwhelm ecol- ogists. I develop a conceptual framework to classify connectivity statistics according to the component of connectivity that they quantify (i.e., inter-patch dispersal likeli- hood, route redundancy, route vulnerability or habitat availability) and the level of the habitat network to which they apply (i.e., element, neighbourhood, cluster or network).

This framework provides much needed clarification of the ecological interpretation of network connectivity statistics and will be an invaluable resource for anyone undertak- ing a habitat network connectivity analysis.

1.4.5 Chapter 6: Conclusions and future directions

This dissertation provides novel insights into how spatial and temporal landscape het- erogeneity can be incorporated into reserve design and habitat connectivity assess- ments. In Chapter 6, I summarize my findings and offer new research directions. The natural extension of this research is to focus on how the structure and connectivity of reserve networks is linked to the long-term maintenance of biodiversity and ecosystem services in human-modified landscapes. Chapter 2

Comparing static versus dynamic protected areas in the Qu´ebec boreal forest

2.1 Abstract

Conservation planning is often based on static mapping of species ranges or habitat distributions. Succession and disturbance alter, however, habitat quality and quan- tity through time especially under global climate and land-use change scenarios; hence, static protected areas may not ensure habitat persistence and species survival. Here, I examined the relative merits of static and dynamic (floating) protected areas for the conservation of American marten (Martes americana) habitat in a dynamic boreal forest of Qu´ebec (Canada). Forest dynamics were modeled using a spatially-explicit landscape-disturbance model and protected areas were selected based on the qual- ity and compactness of marten home ranges using the conservation planning software

MARXAN. Static protected areas were fixed in space during 200 year simulations of boreal forest dynamics, while dynamic protected areas were re-located every 50 years

15 Chapter 2. Comparing static versus dynamic reserves 16 to track dynamic habitat. Dynamic protected areas supported more high-quality home ranges through time than static protected areas. The locations of dynamic protected areas were constrained, however, by the highly fragmented forest patterns created through logging and fire in unprotected areas. My findings emphasize the often-overlooked point that if dynamic conservation planning is to be successful in the long term, the landscape matrix quality surrounding protected areas must be managed in such a way that options remain when it comes to re-planning.

2.2 Introduction

Protected areas play a vital role in safeguarding against further losses of existing biodi- versity and wildlife habitat that result from increased human encroachment on natural spaces (Williams et al., 2004). To fulfill this role, protected areas (hereafter PAs) must not only ensure that multiple elements of biodiversity are adequately represented but that they persist in their protective role through time in dynamic landscapes (Mar- gules and Pressey, 2000; Gaston et al., 2006). In the last two decades, numerous site- selection algorithms have been developed to identify sites to be included in one or more

PAs that maximize the amount of extant biodiversity represented given existing eco- nomic constraints. These algorithms typically rely on static distribution patterns of biodiversity surrogates such as species (Csuti et al., 1997), habitat types (Olson and

Dinerstein, 1998), and land cover types (Pressey et al., 1996) to identify, prioritize, and select candidate PAs. However, because distribution patterns of species and habitats change over time due to natural processes (e.g., population dynamics, succession, and disturbances) and human impacts (e.g., land-use and climate change), site-selection al- gorithms based on representation objectives alone do not necessarily select PAs that ensure the long-term biodiversity persistence (Margules et al., 1994; Rodrigues et al.,

2000). Chapter 2. Comparing static versus dynamic reserves 17

Promoting long-term persistence of biodiversity in PAs therefore requires that site- selection algorithms move beyond simply meeting initial representation targets to- wards incorporating information about the spatio-temporal dynamics of populations and landscapes (Cabeza et al., 2004). Efforts to include population dynamics in site- selection algorithms have treated species-persistence objectives either implicitly by pri- oritizing higher quality habitat (Williams and Araujo, 2000) in large, compact, and well-connected PAs (McDonnell et al., 2002) or explicitly by modeling species-specific persistence through population-viability analyses (Carroll et al., 2003) or metapopu- lation models (Moilanen and Cabeza, 2002). Landscape dynamics have only recently been incorporated into site-selection algorithms and have primarily taken the form of an implementation problem in which PAs must be implemented gradually over several time periods due to budget constraints (Pressey and Taffs, 2001; Meir et al., 2004; Oet- ting et al., 2006). These studies incorporate the possibility that unprotected sites may be lost or degraded before the next implementation period due to anthropogenic land- scape changes such as development or resource extraction. The conservation implica- tions of stochastic landscape processes such as succession, natural disturbance regimes, and climate shifts that operate on timescales longer than the typical PA implementa- tion horizon have just begun to enter discussions about PA design and have remained fairly theoretical (Cumming et al., 1996; Bengtsson et al., 2003; Araujo et al., 2004;

Pyke et al., 2005; Williams et al., 2005, but see Pickett and Thompson, 1978). A few studies have proposed, however, methods to incorporate dynamic ecological processes that provide important ecosystem services such as specialist pollinator interactions, hy- drological regimes, trophic interactions, and ecological diversification of species lineages into quantitative conservation targets (Pressey et al., 2003; Chan et al., 2006).

The Qu´ebec boreal forest is a good system to examine the implications of land- scape dynamics for conservation planning because it is perpetually dynamic due to both stochastic natural disturbances and deterministic human activities. Wildfire is Chapter 2. Comparing static versus dynamic reserves 18 the dominant natural disturbance agent and is highly variable in terms of extent, fre- quency, and ecological impact (Johnson, 1992). Logging, the dominant human distur- bance agent, eliminates or substantially reduces the mature portion of these forested landscapes (Bergeron et al., 1999), and this has motivated the creation of a system of PAs to protect mature forest habitat and act as constraints to the logging sched- ule (Baskent and Keles, 2005). Selecting PAs in landscapes driven by highly variable disturbance regimes may be difficult because representative areas may not exist even at very broad scales (i.e., there may not be a smaller part of the region that is similar to the whole region; Cumming et al., 1996). The lack of representational areas in bo- real forests in terms of an appropriate compositional measure has led some researchers to propose an alternative to static PAs, in which representation and persistence crite- ria are re-assessed periodically through time to create a dynamic system of PAs that change locations to track dynamic habitat (Cumming et al., 1996; Bengtsson et al.,

2003). Dynamic protected areas may be able to overcome some of the difficulties pre- sented by highly variable natural systems when planning for conservation by ensuring that adequate habitat is always available. An intuitive illustration of the dynamic PA concept occurred as early as 1983 in the Gifford Pinchot National Forest when a wind- storm damaged nesting sites in six out of forty-five PAs for spotted owls (Strix occiden- talis) resulting in the re-location of one PA and the merging of two others (Ruediger,

1985).

When considering the implementation of dynamic PAs, an important distinction must be made between PAs implemented over broad, regional scales (Carroll et al.,

2003) and those PAs implemented within multi-use forest landscapes subject to log- ging (Baskent and Keles, 2005). The scale of PAs will determine which criteria and processes should be considered during the selection procedure. At a regional scale, PAs may be large enough that they can maintain internal re-colonization sources under the natural disturbance regime (Pickett and Thompson, 1978); hence, dynamic PAs at this Chapter 2. Comparing static versus dynamic reserves 19 scale are unnecessary because species occurrences alone in these PAs should serve as a surrogate for their long-term persistence. Smaller PAs implemented at local scales may also play an important role in biodiversity protection by preserving species or habi- tats that are under-represented in regional-scale PAs, providing a source of propag- ules and offspring for re-colonization of logged areas, and facilitating the movement of biota across managed landscapes (Lindenmayer et al., 2006). Species persistence in these local scale PAs will be highly dependent on landscape dynamics as well as spatio- temporal population dynamics; hence allowing these PAs to be re-located in response to changes in habitat quality and configuration may improve their ability to conserve target species.

Although several conceptual papers have suggested the use of dynamic PAs in bo- real forest ecosystems (Cumming et al., 1996; Bengtsson et al., 2003; Angelstam et al.,

2004), none have discussed the details of their implementation. In this study, I inves- tigate the potential of using dynamic rather than static PAs to maintain old growth habitat within the boreal forest and explicitly describe the steps necessary to identify, prioritize, and select PAs through time. The total area of PAs was fixed between 10% and 12% of the study area to allow for alternative forest uses such as timber harvest and to comply with conventional conservation targets (McNeely, 1993; IUCN, 1994).

Patches in the static PAs were fixed in space throughout the course of a 200-year sim- ulation of boreal forest dynamics, whereas patches in the dynamic PAs were period- ically re-evaluated and re-located in response to changes in forest conditions due to

fire, succession, and logging. The length of the re-planning interval for dynamic PAs was assumed to be constrained by operational costs associated with PA re-location and was therefore conservatively set as 50 years corresponding to half of the harvest- ing age (100 years). PAs were chosen based on habitat requirements of the American marten (Martes americana Rhoads), which is often used as an indicator species for ma- ture boreal forest biodiversity (Thompson, 1991; Buskirk and Powell, 1994) because of Chapter 2. Comparing static versus dynamic reserves 20

its large home range size (maximum home range size of 1000 ha; Chapin et al., 1998;

Potvin et al., 2000) and its degree of habitat specialization (prefers coniferous/mixed

forest greater than 30 years old; Buskirk and Powell, 1994; Thompson, 1994). To se-

lect PAs, I first identified all potential marten home ranges in the study area-based on

the spatial configuration of forest types and ages (James et al., 2005). I then used the

simulated annealing procedure implementation in MARXAN to find two spatially com-

pact patches of high-quality potential marten home ranges to form a PA (Possingham

et al., 2000). Boreal forest dynamics were simulated using an existing spatially-explicit

landscape-scale model (Fall et al., 2004) that includes sub-models for forest succession,

fire, and logging. I hypothesized that dynamic PAs would outperform static PAs be-

cause the marten home ranges within the static PAs may deteriorate through time due

to losses from wildfire. To compare static and dynamic PAs, I assessed the number,

quality, and connectivity of potential marten home ranges remaining within PAs and

across the entire landscape every 50 years over the course of the 200 year simulations.

2.3 Methods

2.3.1 Study area

The Vermillion study area is a boreal mixed-wood forest management unit that covers

approximately 430,000 ha in south-central Qu´ebec (Fig. 2.1). The northern portion of

the study area is a coniferous boreal forest characterized by black spruce (Picea mar-

iana (Mill.) BSP.), balsam fir (Abies balsamea (L.) Mill.), jack pine (Pinus banksiana

Lamb.) and white birch (Betula papyifera Marsh.). The southern portion is a tran- sition zone between the boreal and temperate forest regions characterized by white spruce (Picea glauca (Monech) Voss), balsam fir, and yellow birch (Betula alleghanien- sis Britt.). Lakes and wetlands are distributed throughout the study area to make up approximately 10% of the total area. Stand-replacing fire is the dominant natural Chapter 2. Comparing static versus dynamic reserves 21 disturbance (H´elyet al., 2000) and logging has been the dominant anthropogenic dis- turbance since the 1950s. Data for the Vermillion study were derived from the 1990 decadal forest inventory carried out by Syst`emed’Information FOResti`erepar Tesselle

(SIFORT) in collaboration with the Qu´ebec Ministry of Natural Resources and Abitibi

Consolidated. Spatial data of tree age and species composition were derived from su- pervised classification of aerial photography and are represented in raster format at a resolution of 0.25 ha (50 × 50 m).

2.3.2 Boreal forest landscape dynamics model

I used a spatially-explicit stochastic raster simulation model that integrates succession, growth, yield, and fire regime information to simulate landscape-scale forest dynam- ics. The model was implemented in SELES modeling platform (Fall and Fall, 2001) and consisted of sub-models for fire, succession, and logging that interact by modify- ing the dynamic forest conditions within each 0.25 ha pixel. Fall et al. (2004), Didion et al. (2007), and James et al. (2007) used this model to investigate interactions among forest management strategies and fire. I used this model to simulate boreal forest dy- namics over 200 years with 10 replicates for both the static and dynamic PA scenarios.

The number of replicates and simulation duration was limited by the data processing time required to move between the simulation model and PA selection software.

Description of sub-models

Fire was modeled as a stochastic process based on empirical estimates of mean fire size and fire return interval (Van Wagner, 1978). The mean number of fires per year was a function of the size of the study area, the mean fire size, and the fire return interval.

Mean Number Fires per Year=(Study Area Size)/[(Mean Fire Size)(Fire Return Inter- val)].

Fire sizes were selected randomly from an exponential distribution with a mean of Chapter 2. Comparing static versus dynamic reserves 22

Figure 2.1: Location of the Vermillion study area in the province of Qu´ebec, Canada.

The map of the study area shows the spatial arrangement of habitat types at a resolu- tion of 0.25 ha/cell. Chapter 2. Comparing static versus dynamic reserves 23

1500 ha (Fall et al., 2004). The exponential model is a commonly made assumption for studies of fire in the boreal and boreal mixed-wood forest and necessarily assumes inde- pendence among fire events (Van Wagner, 1978; Johnson and Van Wagner, 1985; Berg- eron et al., 2002; Fall et al., 2004). Fire return interval, the number of years required to burn an area of size comparable to the study area, was estimated to be 250 years to reflect recent changes in climate (Bergeron et al., 2001). Fire ignition occurred ran- domly and fire events spread randomly from the point of ignition to any eight neigh- boring cells until the total area burned reached the pre-selected size. In this statistical

fire model, fires burned independent of forest age or species composition so that the effects of different PA scenarios were not confounded by uncertainty in the burn proba- bilities for different fuel categories. This simplifying assumption emphasizes the effects of stand-replacing fires in boreal forests which are strongly linked to extreme weather events during which fires burn young and old stands alike (Johnson et al., 2001; Lefort et al., 2003). Such fires account for more than 97% of the area burned, although they account for less than 5% of the total number of fires (Stocks et al., 2002).

Forest succession was modeled using probabilistic successional pathways derived as a function of site-specific soil and drainage conditions and previous vegetation types based on long-term, plot-level data on forest transitions in the Vermillion study area

(Fall et al., 2004). Forest stand age was incremented annually until a maximum of 300 years unless either a fire or logging event occurred, which reset stand age to zero. For- est stand composition likewise followed a probabilistic successional trajectory unless reset by a fire or logging event. In the absence of disturbance, cells continue to age to a maximum of 300 years and site-specific maximum volume. They remain in the max- imum age and volume categories until they are either burned or logged, implicitly as- suming that such stands are in a gap-phase old forest state.

To simulate the effects of logging, the logging sub-model used an area-based an- nual allowable cut (AAC) with a minimum harvest age of 100 years and site-specific Chapter 2. Comparing static versus dynamic reserves 24 road accessibility that reflected plausible management scenarios for the mixed-wood boreal region in Qu´ebec. The AAC was set at 3900 ha (1% of the productive forested study area; Fall et al., 2004; Bergeron et al., 2006; Didion et al., 2007) and volume was estimated for each cell based on yield curves from plot data, but the details are not relevant here because the AAC was based on area not volume. Using an area-based logging target allowed for greater control and transparency in the forest management process by providing a common currency (i.e., forest area) for the logging and conser- vation targets and the fire sub-model. Harvest block sizes were chosen randomly from a uniform distribution between 5 ha and 60 ha based on size ranges common in this re- gion. Harvest blocks were placed preferentially within active “operating areas” and in regions that had high wood volume and road access to capture the tendency for sets of harvest blocks to be grouped for operational efficiency. An operating area is repre- sented as a circle of area 25,000 ha on the landscape that is active for approximately 20 years, after which time it is deactivated and new operating areas are randomly created within productive forest. Up to five operating areas are active at any given time during a simulation. By preferentially harvesting within operating areas, harvest blocks that are close in time will also tend to be close in space. Simple road construction was also included in the harvest model to account for changing site accessibility through time.

Maps of existing roads and projected roads were used to identify the initial road net- work and simulate future development of main roads. Areas further than 2 km from active roads were not available for harvesting until road access expanded to encompass them. Harvest blocks were explicitly connected to current and future main roads by straight-line, spur road segments. Future roads were activated once a spur road con- nected to them or when they were required to bridge newly activated segments with the existing road network. This method of modeling road development allowed road access restrictions to influence the spatial harvesting pattern whereas road-building as- sociated with harvesting reduced access constraints over time. Chapter 2. Comparing static versus dynamic reserves 25

2.3.3 Protected areas

PAs were selected to promote the persistence of marten within the boreal forest study area. The area of the PA was fixed at 10-12% (39,200-47,040 ha) of the forested por- tion of the study area and could potentially support between 47 (mostly low-quality habitat requiring larger home range sizes) and 94 (mostly high-quality habitat result- ing in smaller home-range sizes) marten depending on the density of habitat (percent- age cover) within their home ranges. Each PA system was made up of two patches to increase the chance that some portion of the PA would survive in the event of a catas- trophic fire and could serve as a source of re-colonization (Shafer, 2001). Without im- posing a 2-patch constraint on the number of patches, the PA system would have be- come highly fragmented and scattered across the landscape in very small patches each comprising only one or a few marten home ranges. This type of PA system would be operationally unrealistic to implement. Furthermore, aggregated PA systems that may include lower-quality habitats have been shown to be biologically more valuable than fragmented ones given high levels of uncertainty concerning the effects of fragmentation on species persistence (Moilanen and Wintle, 2006).

Static and dynamic protected area scenarios

For the static PA scenario, both patches of the PA were selected with MARXAN once at the beginning of the 200 year simulation. This static PA was used as the initial PA for the other PA scenarios. For the dynamic PA scenario, both patches were reselected with MARXAN every 50 years during the 200-year simulation (Table 1). I also ex- plored a hybrid of these two scenarios, the static-core scenario, where one patch of the

PA was fixed in space for the duration of the simulation and the other was reselected every 50 years. Two baseline scenarios were simulated for reference: a no-logging sce- nario in which only succession and fire were simulated to estimate the historical range of natural variability; and a no-PA scenario in which succession, fire, and logging were Chapter 2. Comparing static versus dynamic reserves 26

Table 2.1: Description of alternative protected area (PA) scenarios.

Scenario Protected Area Logging Example

(Year 0, 50, 100, 150, 200)

No PA None Yes

Static PA 2 static patches Yes

1 static “core” patch Static-core PA Yes 1 dynamic patch

Dynamic PA 2 dynamic patches Yes

No Logging Entire study area None

In the example maps, black represents protected areas (areas where logging is excluded)

and white represents areas that are potentially subject to logging. Succession and fire

occur across the entire landscape. Chapter 2. Comparing static versus dynamic reserves 27 simulated without designating a PA. Each of the five scenarios was replicated 10 times using the boreal forest simulation model.

Protected area selection method

The PA selection process involved two steps. First, the study area was divided into planning units corresponding to nonoverlapping, potential marten home ranges based on the spatial configuration of preferred habitat (Fig. 2.2a; James et al., 2005). Plan- ning units must be non-overlapping in order to divide up the study area into poly- gons for potential selection by the reserve-selection algorithm. Marten are intrasexu- ally territorial although male home ranges generally overlap with one or more female home ranges depending on habitat conditions and marten densities (Payer et al., 2004); therefore, using non-overlapping home ranges provides us with a conservative estimate of marten population size. This home-range delineation method assumes that an indi- vidual marten may occupy a home range of less than the maximum size provided there is a sufficient density of preferred habitat within that home range (Schultz and Joyce,

1992; Thompson, 2004). Preferred marten habitat is mature coniferous and mixed for- est (Buskirk and Powell, 1994; Thompson and Harstad, 1994) and at least 50% of each marten home range should consist of this preferred habitat type (Schultz and Joyce,

1992; Potvin et al., 2000). Hence, with a maximum marten home-range size of 1000 ha

(Potvin et al., 2000), any area up to 1000 ha that contained at least 500 ha of habi- tat was considered a viable home range. To identify home ranges with a high density of habitat within them, an expanding moving window centered at each cell was ap- plied iteratively across a raster layer of marten preferred habitat. The resulting map of the habitat densities within home ranges (Fig. 2.1a) was used to prioritize planning units (home ranges) for inclusion in the PA. I used this relatively coarse method of home-range delineation to reduce the analysis time and to make best use of the habi- tat information tracked in the boreal forest simulation model (forest age and com- Chapter 2. Comparing static versus dynamic reserves 28 position) while keeping assumptions about space-use and life history at a minimum.

This method allows us to go one step beyond simply considering the amount of habitat available to marten in PAs by also incorporating spatial aspects of habitat for species

(i.e. the spatial juxtaposition of preferred and non-preferred habitat types) and re-

flected the variability of marten home range sizes (reviewed in Buskirk and McDonald,

1989).

The second step involved selecting planning units to comprise the PA. I formulated the PA selection problem as a “minimum set covering problem” (Cabeza and Moila- nen, 2001): find the set of planning units that minimizes the inclusion of low-density home ranges within a fixed area PA. Formulating the PA selection problem in this way allowed us to employ the existing site-selection software MARXAN (version 1.8.2; Ball and Possingham, 2000; Possingham et al., 2000) which uses simulated annealing to ap- proximate a global optimum for large minimum set covering problems. MARXAN has traditionally been used to select a PA that minimizes the area of the conservation area while meeting fixed targets for the biodiversity surrogate (McDonnell et al., 2002). I simply fixed the target area instead and minimized the difference between the max- imum value of the biodiversity surrogate and the actual value in each planning unit

(i.e. 100% actual density of habitat in home range). This was done by modifying the

MARXAN input files: the conservation-feature file was modified to describe the area- based conservation target for the PA (47,040 ha); the conservation feature distribution

file was modified to describe the size of each planning unit; and the planning unit file was modified to describe the difference between the actual density of habitat in the planning unit (home range) and the maximum possible habitat density (100%).

MARXAN solves minimum set covering problems by minimizing a multivariate ob- jective function that gives a value for a collection of planning units as if that collection constituted a PA. Although I modified the objective function, it still consisted of two main sections: the first was a penalty for not meeting the target area (47,040 ha) and Chapter 2. Comparing static versus dynamic reserves 29 the second was a measure of the ‘cost‘ of the PA based on the density of habitat within marten home ranges and the boundary length of the entire PA. The objective function

I implemented was: X X total cost=APF ×AreaPenalty+BLM BoundaryLengths+ (100-HabitatDensity), P Us P Us where PUs are planning units (marten home ranges), APF is the area penalty factor, a term that assigns a penalty for failing to meet the PA area specified and BLM is the boundary length modifier, a weight used to control the spatial aggregation of selected planning units. The AreaPenalty is a penalty for not adequately meeting the target area for the PA; it is proportionate to the difference between the PA area and the tar- get area. The BLM is used to reduce fragmentation of the PA. If the BLM = 0, the algorithm performs without spatial constraints.

Values of the BLM >0 increase the cost of the boundary length and encourage so- lutions that aggregate planning units into fewer, compact clusters with shared bound- aries. I assigned a value of 106 for the APF and adjusted the value for the BLM such that the PA comprised exactly two distinct patches. I used an adaptive annealing sched- ule with one million iterations and 10,000 ‘temperature’ decreases. The temperature is the probability of accepting a change in the set of planning units that comprise the

PA even if this change produces a higher objective function value. This probability de- creases throughout the algorithm once the chance of prematurely finding a local mini- mum is reduced. MARXAN was run 10,000 times and I kept the best solution (i.e. the solution that produced the lowest value for the objective function) as the final PA (Fig.

2.2b).

Comparing protected area scenarios

To compare the PA scenarios (static, static-core, and dynamic) with the baseline sce- narios (no-logging and no-PA), I performed 10 replicates of each scenario and moni- tored two characteristics of marten home ranges both within PAs and across the entire Chapter 2. Comparing static versus dynamic reserves 30

Figure 2.2: Illustration of the two steps of the PA selection process. The first step was to convert potential marten habitat, based on forest composition and age, into potential marten home ranges, based on the spatial configuration of potential habitat and marten movement abilities. The second step was to select potential marten home ranges for inclusion in the PA using the PA selection software MARXAN. MARXAN selected home ranges based on the density of the habitat within them and the bound- ary length of the PA. Home ranges with high habitat density that were compact were favored for inclusion in the final PA. Chapter 2. Comparing static versus dynamic reserves 31 landscape every 50 years: the number of home ranges and the mean density of habi- tat within each home range (home-range habitat density). Within PAs I also moni- tored both the amount of habitat and the amount of area covered by home ranges as a percentage of the area of the PA. To measure home-range properties within static

PAs (and the core patch of the static-core PAs), I re-ran the home-range delineation algorithm strictly within the boundaries of protection. I tracked the boundary length- to-area ratio of PAs through time using the corrected perimeter-area ratio (Baker and

Cai, 1992). This ratio corrects for polygon size by dividing the perimeter of a polygon by the square root of the product of 4π and the area of the polygon. The corrected perimeter-area ratio for a PA was computed as the average of the perimeter-area ratios for both patches within the PA. Its value is always greater than or equal to 1 (with a value of 1.0 for perfect circles) and becomes arbitrarily large for long, skinny polygons.

At the landscape-scale, I also monitored home-range connectivity (the connected- ness among home ranges) measured as the density-weighted minimum planar graph

(DWMPG; O’Brien et al., 2006). This type of graph-theoretic connectivity analysis summarizes the spatial relationships between landscape elements by building a “graph”

(Harary, 1969) consisting of a set of “nodes” that represent landscape elements and

“links” that represent the potential movement of an organism among them (Urban and Keitt, 2001). In this particular application, the nodes were marten home ranges and the links were Euclidean distances among them. DWMPG is based on a partic- ular type of graph called a minimum planar graph which consists of the maximum number of noncrossing links that connect all nodes (O’Brien et al., 2006). To calcu- late DWMPG, link lengths are weighted by the density of habitat in the pair of home ranges that they connect and then a mean length is computed for all links in the land- scape. Mean DWMPG measures both the structural connectivity and the habitat den- sity (quality) of connected home ranges. Its value is large when home ranges are far apart and have low habitat densities, whereas it is small when home ranges are close Chapter 2. Comparing static versus dynamic reserves 32 together and have high habitat densities.

2.4 Results

None of the PA scenarios (static, static-core, and dynamic) dramatically outperformed the others based on any of the measured properties of marten home ranges across the entire landscape over the 200 years of simulation: number (Fig. 2.3a); habitat density

(Fig. 2.3b); or connectivity (Fig. 2.3c). Error bars representing ± two standard errors were highly overlapping among the three scenarios at this landscape-scale (Fig. 2.3).

However, the dynamic and static-core PA scenarios performed at least as well, if not slightly better than the static PA scenario at all time steps for all home-range prop- erties. Not surprisingly, all three PA scenarios were always worse than the no-logging scenario and better than the no-PA scenario. The number of home ranges across the entire landscape decreased initially but then remained fairly constant through time for all PA scenarios (Fig. 2.3a). The density of habitat within home ranges across the entire landscape also remained fairly constant for all scenarios except for the no- logging scenario, which improved gradually, and the no-PA scenario, which declined

(Fig. 2.3b). The dynamic PA scenario produced the most consistent home-range habi- tat density across the landscape through time; however, even for the static PA sce- nario, mean home-range habitat density only ranged from 67.88% to 72.80%. The con- nectedness among all home ranges in the landscape decreased for all three PA scenar- ios, as illustrated by the increase in mean DWMPG (Fig. 2.3c). The dynamic PAs maintained connectivity better than the static and static-core PAs for the first 150 years of simulation. The dynamic and static-core PAs resulted in the highest levels of connectedness among home ranges after the full 200 years of simulation and differ- ences among the three scenarios were most pronounced at this final time step; however, these differences were still small enough and error bars were large enough that they Chapter 2. Comparing static versus dynamic reserves 33 may have been due to stochasticity alone.

I also tracked home-range properties strictly within the PAs and found stronger evidence of the dynamic and static-core PA scenarios outperforming the static PA scenario (Fig. 2.4). I standardized all PA-scale analyses by the area of the PA, which

fluctuated between 10.34% and 12.13% of the forested portion of the study area. The percentage of PAs that contained habitat was higher in the dynamic PAs than the static-core or static PAs and was more consistent through time (Fig. 2.4a). Differences among scenarios were larger when considering either the percentage of PAs that con- tained home ranges (Fig. 2.4b) or the number of home ranges in PAs (Fig. 2.4c) rather than simply the amount of habitat. The dynamic and static-core PAs maintained sig- nificantly higher percentages of home ranges than the static PAs over time because they were re-selected by MARXAN at the time of analysis whereas the percentage of home ranges in static PAs declined due to stochastic losses from fire (Fig. 2.4b). The number of home ranges in the dynamic PAs was also consistently higher than the static

PAs but the static-core PAs were more variable (Fig. 2.4c). In contrast, habitat den- sity of home ranges within PAs was best maintained by the static scenario; however, after 100 years mean values for all scenarios were very similar and error bars among all scenarios were highly overlapping (Fig. 2.4d). Variability in the habitat density of home ranges was highest in the static-core scenario, which ranged from 58% to 82%, and lowest in the static scenario, which only ranged from 72% to 81%. Taken together,

Figures 2.4b-d show that home ranges in static PAs were less numerous and cumula- tively covered a smaller area of the PAs but had similar, if not slightly higher habitat density than home ranges in the static-core and dynamic PAs.

The boundary length-to-area ratio of the static, static-core, and dynamic PAs showed a different pattern (Fig. 2.4e). The ratio was constant for the static PA because both patches in the static PA were fixed in space. Similarly, the static-core PA scenario maintained a consistent ratio largely because the core patch (which was fixed in space) Chapter 2. Comparing static versus dynamic reserves 34

Figure 2.3: Mean properties of all home ranges in the study area (both inside and out- side of the PAs) over 10 replicates for the alternative PA scenarios at initial conditions and at 50, 100, 150 and 200 years of simulation: (a) number of home ranges; (b) mean density of habitat within home ranges and (c) mean density-weighted minimum pla- nar graph (DWMPG). Error bars represent ± two standard errors and are offset to the left (static-core scenario) and right (static scenario) of mean values to facilitate visualization. Chapter 2. Comparing static versus dynamic reserves 35 was very compact and nearly circular. In contrast, the ratio sharply increased for the dynamic PAs, in the final time step suggesting that high quality home ranges became increasingly discontinuous in the dynamic PA scenario.

2.5 Discussion

In this paper I compared the performance of static and dynamic PAs in a dynamic forest landscape based on their abilities to maintain through time important habitat characteristics of a model species, the American Marten, as it is an indicator of ma- ture forest. Dynamic PAs maintained more home ranges than static PAs when im- plemented within the spatially-explicit boreal forest dynamics model with a clear-cut logging regime and an annual allowable cut of 1%. Home ranges within dynamic PAs were of comparable habitat density to home ranges in static PAs even though they were more numerous. The benefits of dynamic PAs were only observable when home- range characteristics were considered within PAs and were not seen at the landscape- scale. All three PA scenarios were much more similar to the no-PA scenario than the no-logging scenario at the landscape-scale likely due to the area constraint on PAs of

10-12% of the forested area. These general patterns can be attributed to the PA selec- tion methods because the annual allowable cut of the logging regime was kept constant in all scenarios (i.e. the logging rate was not decreased when a portion of the area was protected). The performance of dynamic PAs was constrained by the fragmentation of forest in unprotected areas due to interactions between fire and harvest, which limited options for re-locating PAs during the 200 year simulations. The increasing boundary length-to-area ratio of the dynamic PAs through time reflected a reduction in the con- tiguity of high-quality home ranges remaining on the landscape.

Several conceptual papers have suggested the use of dynamic PAs in boreal forest ecosystems (Cumming et al., 1996; Bengtsson et al., 2003; Angelstam et al., 2004); Chapter 2. Comparing static versus dynamic reserves 36

Figure 2.4: Mean properties of PAs over 10 replicates for the static, static-core, and dynamic PA scenarios at initial conditions and at 50, 100, 150 and 200 years of simu- lation: (a) percentage of PAs that is habitat (i.e. the area of habitat within each PA divided by the total area of the PA); (b) Percentage of PAs identified as home ranges

(i.e. the cumulative area of home ranges within each PA divided by the total area of the PA); (c) Standardized number of home ranges (i.e. the number of home ranges within each PA divided by the maximum possible home ranges of 100% habitat density that would fit in the area of the PA); (d) Mean density of habitat within home ranges in the PAs and (e) Corrected boundary length-to-area ratio of PAs (see text for de- tails). Error bars represent ± two standard errors and are offset to the left (static-core scenario) and right (static scenario) of mean values to facilitate visualization. Chapter 2. Comparing static versus dynamic reserves 37 however, they have not discussed the details of their implementation. My study is the

first to have fully implemented a dynamic PA by selecting PAs every 50 years with

MARXAN and simulating landscape change in a boreal forest with a spatially-explicit stochastic landscape model. My results are undoubtedly influenced by the way in which

I defined focal species habitat, implemented the dynamic PAs, represented landscape processes and captured interactions between the disturbance regime and PA identifi- cation. Two critical implementation parameters likely affected the performance of dy- namic PAs: the length of the re-planning interval and the logging regime applied in unprotected areas. The longer the interval between re-planning, the larger effect log- ging will have on unprotected areas which will, in turn, limit continuous tracks of for- est that could serve as options for re-locating PAs. However, if the replanning interval is too short, then the probability of a PA burning would be too small to merit the ef- fort required to re-plan and re-locate the PA. Therefore, an intermediate replanning interval informed by both the logging and fire regimes is desirable.

This problem of choosing a re-planning interval is also of interest to studies exam- ining the periodic acquisition of PAs over an implementation horizon comprising sev- eral time periods (Costello and Polasky, 2004; Meir et al., 2004; Snyder et al., 2004;

Drechsler, 2005; Rush O’Hanley et al., 2007). These studies recognize that budget con- straints force PAs to be implemented over several planning periods and that uncer- tainty in site availability and habitat loss during the implementation horizon may com- promise the effectiveness and efficiency of the PAs selected. They have considered a range in the number of planning periods from two to ten (Costello and Polasky, 2004;

Snyder et al., 2004; Drechsler, 2005; Strange et al., 2006) without specifying the length of each planning period (with the exception of Meir et al., 2004, who recommended an annual re-planning interval over a 10 year planning horizon). Although the length of the planning periods would affect the probability of site availability in these studies, they have placed greater concern on the number of planning periods because adding Chapter 2. Comparing static versus dynamic reserves 38 more planning periods increases the computational load. Oetting et al. (2006) imple- mented a dynamic PA selection process by updating their broad-scale land acquisi- tion plan in Florida at six-month intervals during a 10 year implementation horizon in response to land-use changes and conservation actions. Strange et al. (2006) showed that allowing planners to swap PAs that become degraded through time increases the overall efficiency of the PA system in terms of the number of species represented for a given budget. However, they did not specify the length of the re-planning interval and they assumed that degradation of PAs was irreversible and that PA selection was not spatially-explicit. Given that I simulated boreal forest dynamics over 200 years and selected spatially-explicit PAs using traditional single-planning-period site-selection methods, the planning periods used in these studies do not directly apply. In this sys- tem, a “dynamic” re-planning period may be appropriate, in which re-planning only occurs when the conservation value of a PA has dropped below some threshold (e.g., due to fires). I identify the issue of re-planning intervals as an area that needs more re- search, particularly as the dynamics of ecological and human systems are increasingly acknowledged as important drivers for the persistence of biodiversity in PAs.

Another implementation parameter that likely affected the performance of the dy- namic PAs was the type of logging regimes implemented in unprotected areas. It is becoming increasingly recognized that conservation of forest biodiversity is not only de- pendent upon protected PAs, but also the management of the forest matrix surround- ing them (Lindenmayer et al., 2006). Studies examining the consequences of destroy- ing habitat that is not selected for inclusion in a PA have shown that the probability of occurrence of the species inside the PA is reduced when all habitat outside PAs is lost (Cabeza, 2003; Cabeza and Moilanen, 2003). In the case of multi-use boreal forests however, habitat outside PAs is not completely lost due to the arrangement of logging activities in space and time. The spatial and temporal distribution of logging can be managed to promote mature forest habitat retention in the unprotected areas of the Chapter 2. Comparing static versus dynamic reserves 39 landscape (Lindenmayer et al., 2006). Designing alternative logging regimes may in- volve modifying three variables (Bergeron et al., 1999): size of the harvest unit, inter- nal composition of the harvest unit, or period of time between successive logging events in a harvest unit (i.e., rotation length). Imposing a spatial constraint on the minimum- harvest-unit size would help to reduce fine-scale fragmentation and if harvesting oc- curred in patterns more in line with marten home ranges, then it may be easier to find new, good protected areas during re-planning. Modifying the composition of harvest units under the clearcutting system of even-aged stand management would help main- tain more structural complexity in the form of snags and coarse woody debris within harvest units. Yet, the internal structural heterogeneity of harvest units would not af- fect my assessments of dynamic PAs as it is at a finer scale than I have modeled here.

The rotation length determines the rate at which the landscape is disturbed and consequently the proportion of mature forest on the landscape. Longer rotation lengths would increase the options for re-location of the PAs to mature forest areas; however, it would also reduce the volume of timber flowing from the forest (Bergeron et al., 2001).

Another option is to protect a training PA that is a large tract of land that will be suitable mature forest habitat during the next re-planning period even if it is not suit- able during the current planning interval. This option suggests the need to conserve an area larger than “needed” to ensure sustainable “recruitment” of high quality PAs over time. This option may reduce negative effects on the logging industry while increas- ing the probability of a high quality option for re-location of the dynamic PA. Initial conditions in this study were created by real interactions among management and fire and represent a realistic set of spatial constraints with regards to forest age and species composition in this region of the boreal forest. Real options for management in gen- eral, and for PA selection specifically, are necessarily constrained by the history of dis- turbance and management within a study area (Gustafson, 1998) and will also depend on current management and anticipated future changes in disturbance. Although the Chapter 2. Comparing static versus dynamic reserves 40 results of this study are specific to the study region, disturbance regime, and stochastic model used, the general qualitative messages emerging from these analyses are broadly applicable and should be instructive for many different systems. One important mes- sage that can be extracted from this localized case study is that small, dynamic PAs will not necessarily solve the problem of our current unwillingness to set aside large

PAs that can accommodate landscape dynamics within their boundaries (Pickett and

Thompson, 1978). The ability of dynamic PAs to safeguard against further losses of biodiversity will undoubtedly hinge on the details of their implementation and a key component of their success will be to manage the matrix surrounding PAs in such a way that options remain when it comes to re-planning. Chapter 3

Incorporating consumer-resource spatial interactions in reserve design

3.1 Abstract

The persistence of species in reserves depends in large part on the persistence of func- tional ecological interactions. Despite their importance, however, ecological interac- tions have not yet been explicitly incorporated into conservation prioritization meth- ods. I show here a general method for incorporating consumer-resource interactions into spatial reserve design. This method protects spatial consumer-resource interac- tions by protecting areas that maintain the connectivity between the distribution of consumers and resources. I illustrate this method with a conservation planning case study of a mammalian predator, American marten (Martes americana), and its two primary prey species, red-backed vole (Clethrionomys rutilus) and deer mouse (Per- omyscus maniculatus). The conservation goal was to identify a reserve for marten that comprised 12% of a forest management unit in the boreal forest in Qu´ebec, Canada. I compared reserves developed using analysis variants that utilized different levels of in- formation about predator and prey habitat distributions, species-specific connectivity

41 Chapter 3. Consumer-resource interactions in reserves 42 requirements, and interaction connectivity requirements. The inclusion of consumer- resource interactions in reserve selection resulted in spatially aggregated reserves that maintained local habitat quality for the species. This spatial aggregation was not in- duced by applying a qualitative penalty for the boundary length of the reserve, but rather was a direct consequence of modelling the spatial needs of the interacting con- sumer and resources. This method for maintaining connectivity between consumers and their resources within reserves has been made available via public software and can be applied even under the most extreme cases of either complete spatial overlap or complete spatial segregation of consumer-resource distributions.

3.2 Introduction

Nature reserves can be designated for the protection of single species (Carroll and Miquelle,

2006; Xu et al., 2006), multiple species (Pressey, 1999; James et al., 2005; Holzkamper et al., 2006), or whole ecological communities (Noss, 1987; Hunter, 1991; Done and Re- ichelt, 1998; Chan et al., 2006). Ultimately, the persistence of species or communities in these reserves will depend on the maintenance of key ecological interactions such as mutualisms, competition, and trophic interactions (Sinclair and Byrom, 2006; Baskett et al., 2007). The local extirpation of any species in a community can impact these key interactions and could lead to a series of cascading changes through the community in terms of habitat for other species, foodweb structure, and nutrient cycling (de Roos et al., 1998; Green and Sadedin, 2005). Despite their importance however, ecological interactions have been incorporated into multi-species reserve-selection methods only implicitly (e.g., by focusing conservation efforts on strongly interactive species (Soul´e et al., 2003); or by identifying a minimum reserve area that incorporates natural dis- turbance and ecological processes (Leroux et al., 2007b)), even when the need to con- sider interactions is known (Gerber et al., 2003). Data are often sparse or absent alto- Chapter 3. Consumer-resource interactions in reserves 43 gether, but when they are available, the persistence of ecological communities would benefit from the inclusion of key interactions into conservation prioritization.

Of the many types of known species interactions, consumer-resource (including predator-prey and host-parasitoid) interactions are the principal driver of structure and function in many ecological communities (Sih et al., 1998). Protecting consumer- resource interactions requires maintaining areas for consumers and their resources to encounter one another (Van de Koppel et al., 2005). The spatial co-occurrence of con- sumers and resources is influenced by the spatial configurations of both consumers and resources, limitations on their spatial movements (Van de Koppel et al., 2005), and adaptive habitat selection (Abrams, 2007). Protecting only a portion of the spatial distributions of consumers and resources in reserves creates a network of habitat frag- ments that can alter the spatial properties of consumer-resource interactions and may in turn, affect population dynamics and persistence (Hoopes et al., 2005; Ryall and

Fahrig, 2006). Consumer-resource models have been used to derive generalized guide- lines for reserve design (Micheli et al., 2004; Baskett et al., 2007) such as the minimum reserve patch size required to protect consumers and resources (May, 1994; Holt, 1997;

Swihart et al., 2001; Prakash and de Roos, 2002). However, accounting for the spatial co-occurrence of consumers and their resources when selecting the location of reserves remains an open challenge with implications for species persistence in reserve networks.

I present a general method for incorporating spatial consumer-resource interactions into reserve-selection using the spatial conservation prioritization software, Zonation

(Moilanen et al., 2005; Moilanen, 2007; Moilanen et al., 2008). The proposed method prioritizes areas for conservation based on their contribution to the maintenance of high quality habitat and connectivity between the distributions of consumers and re- sources. Consumer-resource interactions are modeled via an interaction kernel which defines the probability distribution of foraging distances based on the movement abil- ities of the consumer. Objective ecological measures of species-specific habitat qual- Chapter 3. Consumer-resource interactions in reserves 44 ity and connectivities at the scale of individual home ranges can easily be adapted to the interaction kernel approach. I illustrate the use of this method with a conservation planning study from a boreal forest management unit in Qu´ebec (Canada) prioritizing areas for inclusion in a reserve based on the distribution of habitat for the American marten (Martes americana) and two of its prey, the Red-backed vole (Clethrionomys rutilus) and the Deer mouse (Peromyscus maniculatus). The habitat requirements of the marten and its prey do not completely coincide, and the home range sizes and population-level connectivities of the species are very different (Table 1), which makes quantitative conservation planning in this region nontrivial.

3.3 Methods

Before presenting a method that incorporates consumer-resource interactions into reserve- selection using Zonation, I briefly summarize the key features of the Zonation reserve- selection method and algorithms. I conclude this section with a description of the con- servation case study used to illustrate the method.

3.3.1 Summary of the Zonation reserve-selection algorithm

Zonation is a spatial conservation planning tool that calculates maps of conservation priority. As input it uses primarily GIS raster maps of distributions of biodiversity fea- tures (here species), with local occurrence levels described by presence/absence, abun- dance, density, or probability of occurrence in each grid cell. Zonation produces a spa- tial priority ranking of the grid cells through the landscape via iterative removal of the least important remaining grid cell (Moilanen et al., 2005). “Least important” is de-

fined as the cell contributing the smallest marginal value to the remaining conservation value, the definition of which includes considerations such as local occurrence levels of species, species priorities (weights), remaining range sizes, complementarity of species Chapter 3. Consumer-resource interactions in reserves 45 composition, connectivity requirements of species and uncertainty of species distribu- tion information (Moilanen et al., 2005; Moilanen, 2007). On the basis of a priori, user- defined thresholds, the top“Z%” of cells (e.g., 12% as suggested by IUCN, 1994) that jointly contribute to the greatest conservation value can then be considered one of sev- eral possible near-optimal solutions for conservation planning.

Each species can be individually weighted in Zonation to reflect its conservation priority status, economic value, taxonomic value, or its past distribution loss (Arpo- nen et al., 2007; Early and Thomas, 2007; Moilanen, 2007; Kremen et al., 2008). Data sets having millions of grid cells can be analysed using Zonation (Kremen et al., 2008;

Moilanen et al., 2008). The user manual (Moilanen and Kujala, 2006; Moilanen et al.,

2008) and software are freely available. In my application, I used the additive-benefit- function formulation (Arponen et al., 2005; Cabeza and Moilanen, 2006; Arponen et al.,

2007; Moilanen, 2007), which incorporates additivity across local occurrence levels and additivity across species when aggregating conservation value (Moilanen, 2007).

When connectivity requirements are not included in the prioritization process, indi- vidual grid cells are treated as effectively independent from each other, which may lead to spatially fragmented reserve structures (Moilanen et al., 2005; Moilanen and Win- tle, 2006). Fragmented reserves are generally undesirable. For example, Gaston et al.

(2002) pointed out negative biological consequences in cells close to the edge of the reserve network (increased disturbance, predation, invasibility, and abiotic changes);

Hanski (1998b) pointed out negative biological consequences due to altered metapop- ulation dynamics; and Groeneveld (2005) pointed out negative economic and logis- tic consequences associated with a large number of small reserve sites. Consequently,

Zonation provides three options for including connectivity requirements into the analy- sis: (1) boundary-length penalty (Possingham et al., 2000; Moilanen and Wintle, 2007);

(2) distribution smoothing (Moilanen et al., 2005; Moilanen and Wintle, 2006; Moila- nen, 2007) and (3) neighborhood-quality penalty (Moilanen and Wintle, 2007), which Chapter 3. Consumer-resource interactions in reserves 46 c c Mixed c 60 60 Mixed Deciduous Coniferous a,b ha 2 ha 2 ha a,b d a,b 0.0010 0.025 0.025 2000m3162m 80m 141m 80m 141m 22136m 990m 990m 0.000090 0.0020 0.0020 Martes americana Peromyscus maniculatus Clethrionomys rutilus HRsize √ HRLinearDimension = × e HRradius MedianDispersalDist =7 e =2/ f =2/ HRsize/π f Table 3.1: Focal species’ habitat, home range, and dispersal parameter estimates. p Thompson and Harstad (1994). Potvin et al. (2000). Buskirk and Powell (1994). Bowman et al. (2001). Bowman et al. (2002). Moilanen et al. (2005). Home range linear dimension Median dispersal distance Home range scale alpha Population scale alpha Parameter Minimum stand agePreferred habitatSecondary habitatHome range sizeHome range radius= a 30 b Coniferous c Mixed d 1000 e f Chapter 3. Consumer-resource interactions in reserves 47

can also operate in a directed fashion in freshwater/riverine systems (Moilanen et al.,

2008). Of these, the latter two options allow inclusion of species-specific connectivity

requirements into conservation prioritization.

In this work, I employed the distribution-smoothing option, which converts distri-

bution maps into connectivity layers via application of species-specific dispersal kernels,

which is an approach structurally identical to that commonly used in the context of

metapopulation modelling (Moilanen and Nieminen, 2002; Moilanen et al., 2005). The

species-specific connectivity, sij, for species j in grid cell i is computed from original occurrence levels (aj) as:

N X sij = exp(−αdin)anj, (3.1) n=1,n6=i where N is the number of cells in the landscape, α is a parameter for the species-specific

dispersal capacity (e.g., Moilanen and Nieminen, 2002), and din is the distance be- tween cells i and n. Dispersal capacity can refer to different scales of movement behav-

ior, from routine movements associated with daily resource exploitation such as food

searching within a home range, to special movements associated with directed displace-

ment events such as interpopulation movements or migration on a regional scale (Ims,

1995; Van Dyck and Baguette, 2005). Therefore α can be scaled accordingly to either

a home-range scale or an inter-population scale to produce connectivity layers describ-

ing either fine- or coarse-scale connectivity requirements (Fahrig, 2003). These species-

specific connectivity layers can be entered into the same Zonation analysis along with a

species’ habitat distribution maps to model different ecological requirements. If several

species-specific connectivity layers are used (e.g., home-range and inter-population con-

nectivity layers), some attention should be paid to their relative weighting. For exam-

ple, an inter-population connectivity layer could be assigned a relatively lower weight

to indicate that it is less important than home-range level connectivity or local habitat

quality. Chapter 3. Consumer-resource interactions in reserves 48

Figure 3.1: Example of the two types of interaction layers used as input for the Zona- tion reserve-selection algorithm: a resource-use intensity layer for resource j (Rj; upper panel) and a resource connectivity layer for consumer k (Ck; lower panel). The re- source distribution is transformed into a resource-use intensity layer by multiplying the resource-occurrence level of each cell by the connectivity between that cell and the con- sumer population (Eq. (3.2)). The consumer distribution is transformed into a resource connectivity layer by multiplying the consumer representation level in each cell by the connectivity between that cell and the resource distribution (Eq. (3.3)). Connectivity among resources and consumers is determined by the parameter β which models the foraging distances of the consumer. In my illustration, I assume that distributions are binary (presence is grey, absence is white). In the resulting Rj and Ck interaction lay- ers, darker colors represent higher values, as a result of the combination of high local value and high connectivity. Chapter 3. Consumer-resource interactions in reserves 49

The Zonation reserve-selection methods summarized above deal with the require- ments of individual species that are treated as independent from each other. Interac- tions among species have not been directly accounted for. One can easily enter the dis- tributions of both a consumer and its resource into the analysis and overlapping areas will be possibly selected for protection, but only if locations of moderate to high lev- els of co-occurrence do exist. However, if only a small fraction of the landscape can be protected and there are many species in the analysis, it is in no way guaranteed that overlapping areas of consumer and resource will be selected. I next address this poten- tial deficiency by introducing the use of interaction connectivity layers.

3.3.2 Introducing novel spatial consumer-resource interactions

into Zonation

I present a variant of the distribution smoothing technique to compute two different in- teraction connectivity layers between a consumer and its resources (Figs. 3.1 and 3.2):

(1) a resource-use intensity layer for resource j (Rj); and (2) a resource connectivity layer for consumer k (Ck). The resource-use intensity layer, Rj, is a transformation of

the resource’s distribution map based on connectivity to the consumer. Rj describes those parts of the resource distribution that are accessible to the consumer (i.e., re-

source locations within the consumers foraging distance). The resource connectivity

layer, Ck, is a transformation of the consumers distribution based on connectivity to

resources scaled by foraging distances. Ck describes those parts of the consumer dis- tribution that have access to the resource (i.e., consumer locations within foraging dis-

tance from resources). These two interaction connectivity layers can be weighted and

included in conservation prioritization along with the species distributions and species-

specific connectivity layers. Chapter 3. Consumer-resource interactions in reserves 50

The local occurrence levels of resource j and consumer k in grid cell i, are denoted

as rij and cik, respectively (Fig. 3.1). The resource use intensity, Rij, of resource j at cell i by consumer k is defined as the local occurrence level of the resource multiplied by connectivity to the consumer distribution,

N X Rij = rij exp(−βkdin)cnk, (3.2) n=1 in which foraging distances of the consumer are modeled as a negative exponential in-

teraction kernel function; parameter βk is the foraging-distance capacity of consumer

k and din is the distance between cells i and n. Similarly, resource connectivity, Cik, of consumer k at cell i is computed as the local occurrence level of the consumer multi-

plied by connectivity to the resource distribution,

N X Cik = cik exp(−βkdin)rnj, (3.3) n=1

using the same parameter βk to model foraging distances. Thus grid cells with high Rj are locations where both the resource is well represented and a relatively high number of consumers are within foraging distance. Grid cells with high Ck are locations where both the consumer is well represented and a relatively high number of resources are within the foraging distance. Note that the connectivity between consumers and re- sources is calculated from cell to cell so it is desirable to have high-resolution distribu- tion maps relative to the species dispersal abilities. Here I am using distribution maps with 50m × 50m (0.25 ha) resolution and the median dispersal of marten is 22,136m

(Table 1). These generic consumer-resource formulations could also represent a range of species interactions including herbivore-plant interactions, host-parasitoid interac- tions, and predator-prey interactions. Furthermore, functions describing the negative consequences of connectivity for some interacting species (Hof and Flather, 1996; Mi- nor and Urban, 2008) could easily be used to include negative species interactions. Re- gardless of the type of connectivity modeled, the interaction will maintain its general Chapter 3. Consumer-resource interactions in reserves 51

form, which is (local habitat quality of the resource/consumer) × f(connectivity to the

consumer/resource) (Koelle and Vandermeer, 2005).

3.3.3 Case study: predator-prey interaction in the boreal for-

est of Qu´ebec (Canada)

I illustrate this method with a conservation planning case study in a boreal forest man-

agement unit (approximately 430,000 ha) in south-central Qu´ebec focusing on the in-

teraction between the American marten (predator) and two of its prey species, the

red-backed vole and the deer mouse. The objective of the conservation plan was to

select 51,600 ha (12% of the management area IUCN, 1994) that would best protect

American marten as it is considered an old growth indicator species (Thompson, 1991;

Buskirk and Powell, 1994; Rayfield et al., 2008). The case study area is character-

ized by a northern coniferous boreal forest, dominated by black spruce (Picea mari-

ana (Mill.) BSP.), balsam fir (Abies balsamea (L.) Mill.), jack pine (Pinus banksiana

Lamb.) and white birch (Betula papyifera Marsh.), and a southern boreal-temperate transition forest, dominated by white spruce (Picea glauca (Monech) Voss), balsam

fir, and yellow birch (Betula alleghaniensis Britt.). Spatial forest composition data

(stand age and species) were derived from supervised classification of aerial photog- raphy and plot-level ground surveys from the 1990 decadal forest inventory carried out by Syst`emed’Information FOResti`erepar Tesselle (SIFORT) in collaboration with the

Qu´ebec Ministry of Natural Resources and Abitibi Consolidated. Spatial forest data are represented in raster format at a resolution of 0.25 ha; hence, the study area con- sists of approximately 1.3 million grid cells.

To incorporate consumer-resource interactions into Zonation reserve-selection, I converted spatial forest composition data into binary habitat distribution maps (habi- tat and non-habitat) for marten, voles, and mice (see Table 1 for species-specific habi- tat requirements). I obtained estimates for species home-range sizes from the litera- Chapter 3. Consumer-resource interactions in reserves 52

ture and estimated species-specific median dispersal distances using the scaling factor

presented by Bowman et al. (2002, see Table 1). The home-range radius and median-

dispersal distance of each species were used to calculate species-specific estimates of

dispersal ability (α) at home-range and population scales to be used in connectivity calculations (Table 1). The estimate of home-range scale dispersal ability (α) for marten

was also used as the foraging distance capacity estimate (β) to be used for creating in-

teraction layers.

3.3.4 Reserve-selection scenario comparison with different com-

binations of species-specific and interaction connectivity

layers

To evaluate the relevance of including a predator-prey interaction in reserve-selection,

I selected a reserve comprising 12% of the landscape using five different combinations

of predator and prey habitat distribution maps, species-specific connectivity layers, and

interaction connectivity layers (hereafter referred to as Scenarios I-V; Table 2).

In Scenario I, reserves were selected without using Zonation by simply overlay-

ing potential home ranges of marten, mice, and voles to identify those marten home

ranges which had the greatest number of prey (mice and vole) home ranges within

them. Marten home ranges with the highest number of mouse and vole home ranges

making up 12% of the landscape were selected for inclusion in the reserve. Potential

home ranges were delineated by assuming that each individual requires a sufficient den-

sity of habitat within an area smaller than a species-specific maximum home-range size

(James et al., 2005; Rayfield et al., 2008). For example, marten individuals require at

least 50% of their home range to consist of suitable habitat (Schultz and Joyce, 1992;

Potvin et al., 2000) and have a maximum home-range size of 1000 ha (Potvin et al.,

2000). An expanding moving window was passed across the binary habitat map and Chapter 3. Consumer-resource interactions in reserves 53 any area up to 1000 ha that contained at least 500 ha of habitat was identified as a potential home range (with an associated measure of home-range quality - habitat den- sity). This deterministic home-range delineation algorithm was run for each species.

The resulting maps showed the density of habitat within potential home ranges and were overlaid to identify high-quality marten home ranges that contained the most high-quality prey home ranges.

Scenarios II through V used Zonation to identify top 12% fractions of the landscape using different local quality, connectivity and interaction components (Table 2). These scenarios were based on habitat distribution maps and estimates of species’ movement and foraging distances; the estimated home ranges described above were not used be- yond Scenario I. When prey-habitat maps were included (Scenarios II through IV), they were cumulatively down-weighted to 0.5 so that they would be only half as impor- tant as the predator-habitat map. Reserves were targeted towards marten because they are indicators of boreal forest biodiversity (Buskirk and Powell, 1994) and their main prey, Red-backed voles and Deer mice, were down-weighted because they are abundant throughout their ranges. Scenario V did not include prey-habitat distributions or prey- connectivity layers in the reserve-selection, which is reasonable as the mice and voles and their habitat are not particularly endangered. Spatial reserve structures proposed by the five different reserve-selection scenarios were contrasted based on their spatial configuration and habitat composition.

3.4 Results

I present the conservation-area designs generated using different considerations of lo- cal habitat quality and connectivity in Figure 3.2. Candidate solutions were compared both in terms of habitat quality and connectivity. Using local habitat quality alone as an evaluation criterion would ignore the importance of reserve spatial configuration Chapter 3. Consumer-resource interactions in reserves 54 Home range Population scale Form of interaction Connectivity b Relative prey weights indicating the combined weight of all prey habitats. b Prey weight IRC 0.5 R&C Yes Yes Table 3.2: Differences among reserve-selection scenarios. pQHPpQHP 0.5IC N/A 0 C Yes Yes Yes Yes pQ 0.5 N/A No No a mQ mQHP mQHP mQHP Scenario names indicate which components were used during reserve-selection. Z indicates that Zonation was used. IIIZ IV Z VZ No. Name IIIZ Home range overlap N/A N/Aa mQHP and pQHP indicate that marten (m) orpopulation prey connectivity (p) (P) quality were of used. habitat IRC (Q), indicates home thatC: N/A range the resource connectivity interaction connectivity) (H) layers were and (R: used. resource-use intensity and In the example maps, black represents protected areas (areas N/A where logging is excluded). Chapter 3. Consumer-resource interactions in reserves 55 and could over-estimate the contribution of reserve patches with high local quality but which are too small and isolated for the marten to occupy (e.g., marten have been shown to use patches of residual forest within clearcut areas if the patches were greater than 15 ha; Buskirk and Ruggiero, 1994). A solution that has apparently high co- occurrence of marten and prey would not be ecologically justified if it is very frag- mented; therefore, I also considered the composition of reserves in terms of the pro- portion of original connectivity they maintained at the scale of marten home ranges, populations, and predator-prey interactions.

The spatial configurations of the top 12% of cells (the reserve size) selected by the

five scenarios show general consensus on important locations to protect: three central patches, a patch in the northwest, and a patch in the south-east (Fig. 3.2a). Increasing the percentage of cells included in the reserve (e.g., 12% increased to 20%) would result in larger and more contiguous reserve patches that would eventually become connected based on proximity (results not shown). Even so, there are notable differences in the aggregation levels and details of the alternative scenarios; importantly, realistic inclu- sion of connectivity effects tends to result in much more aggregated reserve areas. Sce- nario I differed from all others in that the planning units it used for reserve-selection were potential marten home ranges rather than individual grid cells; hence, it did not exhibit any fine-scale fragmentation below the scale of individual marten home ranges

(Fig. 3.2a). Nevertheless, it achieved less home-range and population-level connectivity than did the Zonation analyses that included connectivity requirements quantitatively within the prioritization process.

When species-specific connectivity layers were used in reserve-selection (Scenarios

III, IV and V), selected habitat was very aggregated compared to when selection was based on local quality alone (Scenario II). Including the interaction-connectivity lay- ers resulted in more contiguous reserve patches (Scenarios IV and V compared to Sce- narios II and III). Scenario V produced the most contiguous reserve patches because Chapter 3. Consumer-resource interactions in reserves 56 it included three types of connectivity for the marten, but did not give weight to the

finely fragmented prey species’ distributions, naturally leading to some loss of prey oc- currence levels inside the reserves.

The composition of habitat types within reserves was altered by the inclusion of the predator-prey interaction during reserve-selection (Fig. 3.2b). The proportion of the original distribution of habitat locally suitable for the predator and one or more prey decreased slightly whereas the interaction connectivity increased significantly.

This means that areas are selected where marten and prey occur in close proximity, although not necessarily in the same grid cells. Note that this increased interaction connectivity was observed in this case study despite a relatively high degree of spatial overlap among the habitat requirements of marten, mice, and voles (see Figure 3.3 for other cases of predator-prey spatial habitat matching). The degree of marten home- range connectivity and marten population connectivity in reserves did not notably dif- fer among scenarios regardless of whether interactions were included (Fig. 3.2b), mean- ing that an increase in interaction connectivity could be obtained practically for very small losses of species-specific connectivity.

3.5 Discussion

In this paper I have showed a generalized method for incorporating consumer-resource interactions into spatial reserve-selection. This work was motivated by population- and community-level studies demonstrating that the persistence of single species and entire ecological communities depends on ecological processes that involve not only multi- ple species but also their interactions (Abrams, 1987; Carpenter et al., 1994; McCann et al., 1998; Polis, 1998; Klug et al., 2000; Sabo, 2008). Spatially explicit formulations of these interactions have been promoted in ecological models (Wu and Marceau, 2002).

Including information about the limited mobility of individuals in consumer-resource Chapter 3. Consumer-resource interactions in reserves 57

Figure 3.2: Reserves selected under five different reserve-selection scenarios using dif- ferent combinations of species-specific and interaction-connectivity layers (Scenarios

I-V in Table 2). (a) Areas with the highest conservation value making up 12% of the landscape are shown in black and potential marten habitat is shown in grey in (I). (b)

Habitat composition of reserves was measured as the proportion of original distribu- tion of different habitat types derived from the preferred habitat of martens, mice, and voles (see Table 1). Habitat supporting marten alone was coniferous or mixed forest between 30 and 60 years. Habitat supporting marten and at least one prey was mixed or coniferous forest greater than 60 years. Habitat supporting marten and two prey was mixed forest greater than 60 years. Other compositional measures of the reserves included the proportion of original connectivity among marten and its prey species, marten home ranges, and marten populations. Chapter 3. Consumer-resource interactions in reserves 58 models has produced more stable consumer-resource dynamics than aspatial equiv- alents (Cuddington and Yodzis, 2000); therefore, it behooves conservation planners to incorporate spatially explicit interactions into conservation management decisions.

However, parameterizing Lotka-Volterra reaction-diffusion models (Svirezhev, 2008) or individual-based simulation models (Breckling et al., 2006) by experimentally track- ing individuals remains time consuming, expensive and a difficult challenge (Koenig et al., 1996; Tischendorf, 1997). Until now, the pattern of interactions among species has mainly contributed towards deciding which suites of mutually dependent species should be prioritized for conservation (Witting et al., 2000; Soul´eet al., 2003, 2005;

Van der Heide et al., 2005; Moulton et al., 2007) rather than informing decisions about which locations should be protected to preserve these species. The spatial conservation prioritization method incorporates species-specific habitat-connectivity requirements and inter-specific connectivity requirements between predator and prey habitats.

It is a relevant question whether reserves selected with and without the consumer- resource interaction differ in terms of species representations and spatial configurations.

The results from the reserve-selection scenario comparison show very minor tradeoffs between predator-prey co-occurrence within reserves and contiguity of reserve patches when predator-prey interactions are incorporated (Scenario III compared to Scenario

IV; Fig. 3.2). In fact, when reserves were selected based only on marten habitat re- quirements and the predator-prey interaction connectivity layer (Scenario V), both species co-occurrence and reserve patch contiguity increased (Fig. 3.2). The increased contiguity achieved by including species interactions in reserve-selection makes these reserves more likely to be feasible from a practitioner’s perspective (Possingham et al.,

2000; Gaston et al., 2002; Moilanen and Wintle, 2006, 2007). It is encouraging that such solutions have been suggested based on a quantitative analysis that includes species- specific connectivity considerations - I have used no qualitative method, such as the commonly used boundary length penalty (Possingham et al., 2000), to enforce struc- Chapter 3. Consumer-resource interactions in reserves 59

Figure 3.3: Three cases illustrating varying degrees of co-occurrence between consumer and resource: Case 1, no co-occurrence; Case 2, partial co-occurrence; and Case 3, complete co-occurrence. The corresponding resource-use intensity layer and resource- connectivity layer are shown with darker areas indicating higher connectivity between the consumer and resource. Chapter 3. Consumer-resource interactions in reserves 60 tural aggregation. And of course, a reserve network selected using connectivity and interaction components then by definition represents those components in the chosen areas, indirectly promoting species persistence in the reserve network.

Another relevant question is: did the proposed analysis produce a result which is likely to be better than the result obtained by the simple home-range overlay method of Scenario I? Although I have no reliable spatial population viability analysis (PVA) model available for rigorously evaluating the proposed reserve structures using a com- mon currency, such as extinction risk, I can make the following observations. First, I was able to find a solution (Scenario IV) which was superior to Scenario I in all aspects of co-occurrence of marten and prey, home range, and population level connectivity for the marten and interaction between the marten and prey (Fig. 3.2). Second, whereas

Scenario V is comparable to Scenario I in terms of marten areas with or without prey,

Scenario V is far superior to Scenario I in terms of connectivity components and fea- sibility of implementation (as seen in the solution maps in Fig. 3.2). In Scenario V,

87% of the proposed reserve occurs as spatially contiguous areas that would support at least nine marten home ranges (9000 ha). For comparison, in Scenario I, only 37% of the proposed reserve occurs in continuous areas greater than 9000 ha. Furthermore, in

Scenario V the four largest patches include 95% of the area and the boundary length- to-area ratio of the reserve is low (BL/A = 0.11) compared to that of Scenario I, where the four largest patches include 64% of the area and BL/A = 0.22. Thus, Scenario V appears to be superior to Scenario I both in terms of expected regional population vi- ability and implementability. In summary, it appears that the proposed method for incorporating consumer-resource connectivity has indeed been able to find solutions that are likely to improve on Scenario I both with respect to conservation value and practicality of implementation (Scenarios IV and V).

The large home-range size and high dispersal ability of the marten in this case study smoothed out the fine-scale distribution of the mice and voles when their habi- Chapter 3. Consumer-resource interactions in reserves 61 tat maps were included (Scenario IV). This prioritization of more contiguous reserve patches when predator-prey interactions are included in reserve-selection depends on the degree of similarity between predator- and prey-habitat requirements and disper- sal abilities (Van de Koppel et al., 2005) as well as the initial distribution patterns of each. Modelling efforts to determine the minimum reserve size required to protect predators and prey have also emphasized the importance of including parameters from both predators and prey, regardless of whether the predator is a specialist or generalist

(May, 1994; Holt, 1997; Swihart et al., 2001; Prakash and de Roos, 2002). My results highlight the importance of accounting for predation not only when setting minimum reserve patch sizes but also when deciding on reserve patch configuration.

Although I illustrated the method with a simple predator-prey case study, it can be applied to very large data sets containing hundreds of habitat-quality, connectivity and interaction layers in multi-million element landscapes. Adding more species will inevitably complicate data acquisition and interpretation of planning outputs; there- fore, I modelled the consumer-resource interaction in a simplified form to minimize data requirements and facilitate interpretation. Areas with the highest probability of successful consumption were identified based on the probability of occurrence of con- sumers and resources and their spatial connectivity. I did not explicitly model numeric responses of consumers and resources that would have involved assumptions about den- sity dependence (Krebs, 2002), functional responses (de Roos et al., 1998), or adaptive dynamics (Abrams, 1999). My main assumption was that consumers preferentially uti- lize regions close to their current location for acquiring resources and I modelled this assumption with an exponential-decay curve describing the connectivity between con- sumers and resources. This assumption treats the matrix surrounding habitat patches as homogeneous when proximity is based on Euclidean distance; however, proximity could also be measured in terms of effective cost distance to incorporate information about dispersal barriers through a heterogeneous landscape matrix (Wiens, 1989; With, Chapter 3. Consumer-resource interactions in reserves 62

1994; Adriaensen et al., 2003).

3.6 Conclusion

A novel method for the inclusion of species interactions in reserve-selection was pre- sented. Using a case study, I have demonstrated large influences of connectivity and in- teraction terms on ideal reserve structure. Further evaluation of this method could in- volve trophic models (Baskett et al., 2007) or metacommunity models (Guichard et al.,

2004) to track population sizes of consumers and resources in reserves selected with and without information about the interaction. Empirical evaluations of the reserves selected by this method would be more challenging because the scale of conservation planning for predatory mammals, such as the American marten, is prohibitive to ex- perimental manipulation (Oksanen, 2001). Empirical investigations into the space-use behaviors of the focal predator would strengthen estimates of foraging distance capac- ity and would be of great benefit to this method.

The case study I presented was fairly simple in that it considered only three species with highly overlapping habitat requirements. The predator-prey relationships between marten-vole and marten-mouse allowed us to examine the cases of partial (Case 2, Fig.

3.3) and complete (Case 3, Fig. 3.3) co-occurrence, respectively. Maintaining connec- tivity between consumers and their resources within reserves becomes more challenging as consumer- and resource-habitat requirements diverge; however, this novel algorithm can still be applied under the most extreme case of consumer-resource spatial habi- tat segregation to identify those areas that are most important for the maintenance of the interaction (Fig. 3.3). Furthermore, conservation planning often involves many more species and more interactions (Pressey, 1999). The discrepancies between reserves that include consumer-resource interactions vs. those that do not will likely increase as the number of species and interactions increase and as their co-occurrences decrease. Chapter 3. Consumer-resource interactions in reserves 63

The foraging capacities of the consumers will also affect the degree to which includ- ing the interaction model improves the spatial cohesiveness of the proposed reserve.

Although the consumer in this case study had a large foraging distance capacity, the co-occurrence of the consumer and its resources likely produced a conservative estimate of the effects of including species interactions in reserve-selection. Inclusion of connec- tivity and species interactions in spatial conservation planning should generate reserve networks that are comparatively resistant to extinctions, which is important as human land-use outside reserve borders intensifies (Williams et al., 2004). Chapter 4

The sensitivity of least-cost habitat graphs to relative cost surface values

4.1 Abstract

Maintaining and restoring connectivity among high-quality habitat patches is recog- nized as an important goal for the conservation of animal populations. To provide an efficient measure of potential connectivity pathways in heterogeneous landscapes, least-cost route analysis has been combined with graph-theoretical techniques. In this study I use spatially-explicit least-cost habitat graphs to examine how matrix qual- ity and spatial configuration influence assessments of habitat connectivity. I generated artificial landscapes comprised of three landcover types ranked consistently from low to high quality: inhospitable matrix, hospitable matrix, and habitat. I controlled the area and degree of fragmentation of each landcover in a factorial experiment for a to- tal of 20 combinations replicated 100 times. In each landscape I compared 8 sets of relative landcover qualities (cost values of 1 for habitat, between 1.5 to 150 for hos- pitable matrix, and 3 to 10000 for inhospitable matrix). I found that the spatial loca- tion of least-cost routes was sensitive to differences in relative cost values assigned to

64 Chapter 4. Sensitivity of least-cost habitat graphs 65 landcover types and that the degree of sensitivity depended on the spatial structure of the landscape. Highest sensitivity was found in landscapes with fragmented habitat and between 20 to 50% hospitable matrix; sensitivity decreased as habitat fragmenta- tion decreased and the amount of hospitable matrix increased. As a means of coping with this sensitivity, I propose identifying multiple low-cost routes between pairs of habitat patches that collectively delineate probable movement zones. These probable movement zones account for uncertainty in least-cost routes and may be more robust to variation in landcover cost values.

4.2 Introduction

Faced with the reality of ever increasing habitat fragmentation (Riitters et al., 2000;

FAO, 2006), wildlife conservation efforts have recently focused on protecting and restor- ing habitat connectivity (Crooks and Sanjayan, 2006; Hilty et al., 2006). Habitat con- nectivity is a concept that describes the potential for an animal to move among high- quality habitat patches as a function of both the spatial structure of the landscape and the movement behavior of the animal in response to that spatial structure (Tay- lor et al., 1993; With et al., 1997, 1999; Tischendorf and Fahrig, 2000b; Goodwin and

Fahrig, 2002a,b; Brooks, 2003; Fahrig, 2007). Recent studies have emphasized the roles that the spatial structure and quality of the intervening matrix between high-quality habitat patches has on habitat connectivity (Ricketts, 2001; Tischendorf et al., 2003;

Bowne and Bowers, 2004; O’Brien et al., 2006). Ultimately however, the animal’s move- ment behavior will determine the extent to which inhospitable matrix structure in- hibits movement among habitat fragments (D’Eon et al., 2002; B´elisle,2005). Con- servation planners must therefore describe a landscape from an animal’s perspective in order to understand, measure, and conserve functional habitat connectivity (Wiens,

1989). Chapter 4. Sensitivity of least-cost habitat graphs 66

Animal perceptions of landscape spatial structure are thought to be primarily de- termined by the fitness consequences, such as mortality and reproductive success, ex- perienced by individuals moving through different landcover types (With et al., 1997;

Drielsma et al., 2007; Driezen et al., 2007). Physical resistance (Stevens et al., 2004) and energetic expenditures are two mechanisms that may influence fitness in a given cover type (Graham, 2001; Drielsma et al., 2007). A quantitative description of move- ment behavior within habitats involves assigning organism-specific “cost values” that reflect the quality of the habitat in terms of ecological costs incurred by an individual moving through them. These cost values are often a function of several environmental variables that are relevant for a particular species such as slope, elevation, water, veg- etation cover type and structure, roads, and human settlements (Ferreras, 2001; Cle- venger et al., 2002; Schadt et al., 2002; Chardon et al., 2003; Verbeylen et al., 2003;

Beazley et al., 2005; Cushman et al., 2006; Kautz et al., 2006; Rouget et al., 2006;

Driezen et al., 2007; Epps et al., 2007; Gonzales and Gergel, 2007; LaRue and Nielsen,

2008, see Table 4.1).

Representing a landscape as a cost surface relevant to a particular species involves:

(1) identifying the number of perceptible habitat types in the landscape; (2) ranking all habitat types according to their cost for movement; (3) quantifying the relative cost of landcover types by assigning numerical cost values (note that cost values in some cells may need to be adjusted to account for additional elements such as slope or elevation).

Ideally, cost values should be assessed based on field and experimental data; however, these data are difficult and time consuming to collect, and this has resulted in a large number of studies relying entirely or in large part on expert opinion (Clevenger et al.,

2002; Schadt et al., 2002; Chardon et al., 2003; Verbeylen et al., 2003; Johnson and

Gillingham, 2004; O’Brien et al., 2006; Gonzales and Gergel, 2007). Some studies use

field data on occurrence patterns in order to delineate landcover preferences instead of landcover costs (but see Haddad and Tewksbury, 2005). They employ well-known Chapter 4. Sensitivity of least-cost habitat graphs 67 detected? No Yes Yes (only for very low cost values) Yes Yes “It appeared that withinrange the of resistance values used (1-20), no alarming differences occurred between the accessibility patterns obtained.” pg. 9 Half as many transitions were predicted when a higher relative cost value was assigned to the least-preferred landcover type (pasture). “Only when using verycost low values for matrixthe did connections change, but these values can beto assumed be below arealistic biologically range.” pg. 1479 “All resistance sets we explored had a significantly better fit to theEuclidean data distance.” than pg.Prediction 570 power varied among resistance sets. “the results of thisshow study sensitivity do to variation in these resistance values and allowed us to’most deduct likely’ a resistance set.” pg. 801 Response variable Conclusionbased on simulated dispersers Sensitivity per day among forest remnants under constant energy constraints location of least-cost routes presence/absence in the proximity of source patches presence/absence in the proximity of source patches Range of cost values 1 - 21 Patch accessibility landcover types 9, 11, 6 andrespectively 13 3 1 - 4 Number of transitions 3 1 - 1000 Identification and 6 1 - 100 Predict species’ 2 and 5 1 - 100000 Predict species’ Table 4.1: Chronological andtivity alphabetical studies presentation using of a connec- cost surface to identify least-cost routes. ) ) ) ) Ramphastos Lynx lynx Pararge Sciurus Large and small mammals, birds, and butterflies in the Netherlands Keel-billed toucan ( sulfuratus Eurasian Lynx ( Speckled wood butterfly ( aegeria Red squirrel ( vulgaris et al., 1992) 2001) et al., 2002) et al., 2003) et al., 2003) # Study1 (Knappen Study species Number of Table 4.1 continued on following page 2 (Graham, 3 (Schadt 4 (Chardon 5 (Verbeylen Chapter 4. Sensitivity of least-cost habitat graphs 68 detected? No Yes (only for coarse scale) Yes Yes “There were no discernable differences in the sensitivity analysis least-cost routes that were produced from versions of the costthat surfaces incorporated higher impedance scores for roads and water.” pg. 126 “Hence, error in thevalues cost should only haveminor a influence on the position of key thresholdswhich at large increases inoccur, ECS while the structureconnected of clusters will remain consistent.” pg. 77 “In our example with radiotracking data of hedgehogs, the average z-score of most resistance sets were negative and significantly different from zero, except for twoand sets 12) (11 where thisleast was expected.” also pg. 320 “For...both measures of gene flow, best-fitting models resulted when sloped terrain had 1/20th to 1/10thcost the of movement acrossterrain flat (Fig. 2), withslope the weight of 0.10often most favoured.” pg. 719 Response variable Conclusionleast-cost routes Sensitivity (ECS) of habitat patches in a network. Threshold distances at which large increases in ECS occurred empirical vs. least-cost routes (z-score) Genetic distance and gene flow (population pairwise FST values), Number of migrants per generation (Nm), and Migration rates (M) Range of cost values (1640 was applied to linear dispersal barriers) landcover types 10 and 11 1 - 20 Locations of 6 1.7 - 3.5 Expected cluster size 20 1 - 1000 Cost and location of 3 1 - 100 ) ) ) Ovis ) Puma concolor Rangifer Erinaceus Florida panther ( coryi Woodland caribou ( tarandus caribou Hedgehogs ( europaeus Desert bighorn sheep ( canadensis nelsoni et al., 2006) et al., 2006) et al., 2007) et al., 2007) # Study6 (Kautz Study species Number of Table 4.1 continued on following page 7 (O’Brien 8 (Driezen 9 (Epps Chapter 4. Sensitivity of least-cost habitat graphs 69 detected? Yes “We also found thatthe varying absolute values, and thus range (the incremental increase) of friction values, however, produced greater variation than expected... Finally, our expectation that output from thescenarios nine at the contiguous site would be moreto similar each other thanfrom outputs the fragmented sitesupported.” was pg. 1164 Response variableDistance and Conclusionaccumulated cost values of least-cost routes Sensitivity Range of cost values 0 - 160 (continuous landscape), 0 - 300 (frag- mented landscape) landcover types 18 (continuous landscape), 22 fragmented landscape) ) Sciurus Eastern grey squirrel ( carolinensis and Gergel, 2007) # Study10 (Gonzales Study species Number of Chapter 4. Sensitivity of least-cost habitat graphs 70 methods such as compositional analysis (Aebischer et al., 1993), wherein vegetation cells are ranked according to species’ habitat preferences using a measure of time spent in each landcover type relative to its availability in the landscape (Ferreras, 2001; Gra- ham, 2001; Kautz et al., 2006; O’Brien et al., 2006). Resource-selection indices are also used to quantify landcover preferences by relating environmental variables with oc- currence data through regression models (Boyce and McDonald, 1999; Ricketts, 2001;

Manly et al., 2002). The inverse (Ferreras, 2001; Graham, 2001; Schadt et al., 2002;

Chetkiewicz et al., 2006; Gonzales and Gergel, 2007; LaRue and Nielsen, 2008) or odds ratio (O’Brien et al., 2006) of these measures of landcover preference can then be taken as approximations of landcover cost values.

The resulting cost surface can serve as the basis for analyzing habitat connectivity using a combination of least-cost route analysis and graph-theoretic techniques (Bunn et al., 2000; Chetkiewicz et al., 2006; O’Brien et al., 2006; Theobald, 2006; Fall et al.,

2007). Least-cost route analysis estimates efficient-movement routes and costs between pairs of habitat patches based on the suitability of the intervening matrix (Adriaensen et al., 2003). Graph-theoretic techniques derive holistic measures of habitat connec- tivity by using a dispersal-distance threshold as a surrogate of movement potential among all habitat patches in the landscape (Urban and Keitt, 2001; Fall et al., 2007).

By defining the links of a graph using least-cost routes, spatial information about the habitat patches and the surrounding matrix can be incorporated into graph-based mea- sures of overall habitat graph connectivity. The habitat graph model scales up from pairwise measure of effective distances between patches, which are typically gener- ated by least-cost route algorithms, to consider the connectivity of the entire habitat graph. The graph-theoretic approach also roots measures of habitat connectivity in a conceptual model derived from metapopulation theory whereby the importance of each habitat patch in maintaining overall connectivity of the graph can be attributed to its topological position and source/sink characteristics (Urban et al., 2009). Chapter 4. Sensitivity of least-cost habitat graphs 71

In this study, I addressed the difficulties and ensuing uncertainties that arise when deriving organism-specific descriptions of matrix structure. I examined the sensitivity of habitat-connectivity assessments, based on least-cost links (Adriaensen et al., 2003) and graph-theoretic methods (Urban and Keitt, 2001; Fall et al., 2007), to the way in which we commonly quantify species-specific perceptions of matrix quality. I expected that the spatial location of least-cost links would depend on both the spatial pattern of the habitat and matrix-landcover types as well as the relative cost values of each.

Hence, I ran a factorial experiment to generate artificial landscape spatial patterns and corresponding cost surfaces with three factors controlling landscape spatial pattern and one factor controlling cost values. Comparing habitat connectivity measures among these artificial landscapes allowed us to systematically test the importance of each of the factors.

4.3 Methods

4.3.1 Generation of artificial landscape spatial patterns

I generated landscape spatial patterns in maps of 100 × 100 cells using three land-

cover types: habitat (H), hospitable matrix (HM), and inhospitable matrix (IM). The

maps were generated based on a placement algorithm that arranged landcover types in

fragmented and clumped configurations (Fahrig, 1997, 1998; Tischendorf and Fahrig,

2000a; Tischendorf, 2001; Tischendorf et al., 2003, I follow the notation of Tischendorf

and Fahrig (2000) and Tischendorf et al. (2003) for consistency). The spatial pattern

of each landcover type was controlled by two parameters: COV, which defined the pro-

portion of that landcover type in the landscape; and FRAG, which defined the spatial

arrangement of cells assigned to that landcover type. Defining these parameters for

H and HM constrains the placement of IM; hence the full landscape spatial pattern

can be controlled by the following four spatial pattern parameters: H COV, H FRAG, Chapter 4. Sensitivity of least-cost habitat graphs 72

HM COV, and HM FRAG.

The placement algorithm begins by assigning H to cells of the map. It repeats the

following two steps until the number of cells that have been assigned is equal to the

number of cells determined by H COV: (1) the algorithm randomly selects an unas-

signed cell; and (2) it decides whether or not H will be assigned to this cell based on

the assignment of its eight neighbor cells and the parameter H FRAG. The decision to

assign H to the selected cell is taken if it meets either of the following conditions: (1)

the value of a random number between 0 and 1 is smaller than H FRAG2 or (2) one or

more of the eight neighboring cells already has the assignment H. H FRAG is squared

in order to produce a linear relationship between H FRAG and the number of patches

comprised of cells assigned to H (Figure 1a in Tischendorf and Fahrig, 2000a). When

H FRAG is large, the random number will often be smaller than H FRAG2 and most

of the cells will be assigned to H based on the first condition which will produce a ran-

dom distribution of habitat. When H FRAG is small, most of the cells will be assigned to H based on the second condition which will produce a more contiguous distribution of habitat with larger, less edgy patches. All groups of adjacent cells (based on eight neighbors) with value H are considered as habitat patches and paths are calculated based on all of these patches. The placement algorithm operates in an identical fashion to assign HM to cells and all remaining cells receive the assignment IM (see example landscapes in Fig. 4.1).

I defined the value of H COV as 12% to represent the amount of habitat commonly

targeted for protection in reserves in Canada (Pressey et al., 2003, and references therein).

Evaluating the sensitivity of connectivity assessments, based on least-cost habitat graphs,

in the context of a typical conservation scenario is consistent with the promotion of

this habitat graph approach as a conservation planning tool for ecologists (Urban and

Keitt, 2001; Chetkiewicz et al., 2006; Urban et al., 2009). I manipulated the values of

the three other parameters in a factorial experiment with 20 combinations of the three Chapter 4. Sensitivity of least-cost habitat graphs 73

Figure 4.1: Examples of the simulated landscapes at a resolution of 100 × 100 cells.

Each landscape contains 12% habitat (black squares) and either a) 10% or b) 50% hospitable matrix landcover (HM COV; gray squares). Habitat area is either aggre- gated (H FRAG = 0.05) or more randomly distributed in (H FRAG = 0.5). Hospitable matrix area also either aggregated (HM FRAG = 0.05) or more randomly distributed

(HM FRAG = 0.5) in both landscapes. Chapter 4. Sensitivity of least-cost habitat graphs 74

Table 4.2: Parameters used to generate spatial patterns of landscapes and their fac-

torial combinations. Each cell in the artificial landscape was classified as habitat (H),

hospitable matrix (HM), or inhospitable matrix (IM). The amount and fragmenta-

tion of inhospitable matrix was completely constrained by parameters for habitat and

hospitable matrix. All factorial combinations result in 20 spatial patterns. Spatial

patterns were generated using the (Fahrig, 1997, 1998; Tischendorf and Fahrig, 2000a;

Tischendorf, 2001; Tischendorf et al., 2003).

Factor Description Parameter values No. levels

H COV Amount of habitat (percentage) 12% 1

H FRAG Fragmentation of habitat 0.05, 0.5 2

HM COV Amount of hospitable matrix 10, 20, 30, 50, 70% 5

HM FRAG Fragmentation of hospitable matrix 0.05, 0.5 2

parameters (Table 4.2). I examined H FRAG and HM FRAG values of 0.05 and 0.5

because they produced relatively extreme patterns of randomness and clumpiness re-

spectively for my 100 × 100 cell landscapes. I did not increase the parameters con- trolling the level of fragmentation, H FRAG and HM FRAG, beyond 0.5 to avoid hav-

ing too many single cell patches. To keep this a controlled experiment, I have assumed

that the cell size is the minimum habitat size required; filtering out small habitat patches

would have compromised my ability to tie the results back to the H COV and H FRAG

parameter values. I generated 100 replicates of each given combination of parameters

for a total of 2 000 landscapes.

4.3.2 Cost values to quantify resistance to movement

Cost values were assigned to each of the three landcover types (H, HM, and IM) re-

flecting the ecological costs incurred by an individual of a generic species dispersing Chapter 4. Sensitivity of least-cost habitat graphs 75 through that cover type relative to the costs incurred by dispersing through the pre- ferred habitat (H, which was always assigned a cost of 1). I performed a sensitivity analysis on these cost values to assess their effect on measures of habitat connectivity using least-cost links and graph-theoretic connectivity analyses. The ranking of land- cover types was preserved such that preferred habitat (H) was always least-resistant to movement, followed by the hospitable matrix (HM), followed by the inhospitable matrix (IM; Table 4.3). Cost values were chosen to capture the range presented in the connectivity literature, which reported using fractional, twofold, and orders of mag- nitude differences between landcover costs (Table 4.1). Emphasis was placed on the relative differences of the cost values in different landcover types rather than absolute values (Chardon et al., 2003). I compared eight sets of cost values (C1 to C8) that sys- tematically varied the relative cost between H and HM and between HM and IM. An illustration of the least-cost habitat graphs produced under different sets of cost values is provided in Figure 4.2. I examined four relative cost differences between H and HM

(1.5, 2, 5, and 100 fold) and two relative cost differences between HM and IM (1.5 and

100 fold; Table 4.3).

4.3.3 Graph-theoretic representations of habitat connectivity

using least-cost links

Graph-theoretic connectivity analyses summarize the spatial relationships between landscape elements. This is achieved by building a “graph” (Harary, 1969) consisting of a set of “nodes” that represent preferred habitat patches and “links” that repre- sent the potential movement of an organism among them (Keitt et al., 1997; Calabrese and Fagan, 2004; Treml et al., 2008). When dealing with large landscapes that have many habitat patches and possible links, it is preferable to analyze a subset of all pos- sible links that represents the connected core of the landscape for a given species (Keitt et al., 1997). Chapter 4. Sensitivity of least-cost habitat graphs 76

Table 4.3: Sets of relative cost values used in factorial experiment. H-Habitat, HM-

Hospitable matrix, IM-Inhospitable matrix.

Relative cost values

Cost set H HM IM

C1 1 1.50 2.25

C2 1 1.50 150

C3 1 2 3

C4 1 2 200

C5 1 5 7.50

C6 1 5 500

C7 1 100 150

C8 1 100 10 000

The minimum planar graph is one such graph as it contains the subset of all pos- sible links that includes the maximum number of non-crossing links that are of least weight (therefore loops are allowed; O’Brien et al., 2006; Theobald, 2006). Least-cost routes are only delineated between topological neighbors (i.e., least-cost links) and will never cross habitat patches (see Fig. 4.2). The minimum planar graph allows for computational efficiency while including alternate pathways to reflect important re- dundancy required for population resiliency which is not included in other graph types such as the minimum spanning tree (O’Brien et al., 2006; Fall et al., 2007). In the min- imum planar graph, nodes are 2-dimensional habitat patches that are fixed in space and links connect patches from edge-to-edge rather than from centroid-to-centroid (Fall et al., 2007).

A benefit of the minimum planar graph is that it approximates the complete graph within a reasonable, bounded error (Keil and Gutwin, 1992). Empirical testing on real Chapter 4. Sensitivity of least-cost habitat graphs 77

Figure 4.2: An illustration of a) least-cost links, b) paths, and c) habitat graphs. A link is a route that directly connects two habitat patches (nodes). A path is route that indirectly connects habitat patches along a series of connected links and nodes in which no node is visited more than once. I used least-cost route analysis to identify least-cost links between habitat patches (nodes) to create least-cost habitat graphs.

Least-cost habitat graphs contain 12% habitat (black squares), 10% hospitable matrix landcover (HM COV; light grey squares) and the fragmentation parameter for both is

0.05 (H FRAG, HM FRAG). Habitat patches are connected by least-cost links (dark gray). The sets of cost values differ with respect to the relative costs between the ma- trix landcover types: C1 - Habitat = 1, Hospitable Matrix = 1.5, Inhospitable Matrix

= 2.25; C2 - Habitat = 1, Hospitable Matrix = 1.5, Inhospitable Matrix = 150. Chapter 4. Sensitivity of least-cost habitat graphs 78 landscapes suggests that the mean difference between paths in a complete graph and in the minimum planar graph (Euclidean or least-cost) is much lower than the maximum deviation (and with a space efficiency that can scale to large landscapes; Fall et al.,

2007). Hence, minimum planar graphs could represent an ecological assumption that move through the graph by using stepping-stone paths through topological neighbors (Urban et al., 2009), but this assumption is not necessary because they also approximate a complete graph.

The potential for movement between a pair of high-quality habitat patches (graph link) is based on the assumed movement behavior of the organism. If one assumes the organism follows straight-line movement behavior then links that minimize the geo- graphical (Euclidean) distance of the route traveled between habitat patches will be identified. However, if one assumes that the organism follows least-cost movement behavior then least-cost links that minimize the cumulative cost along the route will be identified (Halpin and Bunn, 2000; O’Brien et al., 2006; Driezen et al., 2007). The length of both Euclidean and least-cost links can be measured in two ways: (1) the ge- ographic length in metric distance units (e.g., meters or kilometers); or (2) the effective length as the cumulative sum of the cost values in cells traversed along the link multi- plied by the length of the link segment crossing those cells.

In this study, I computed the minimum planar graph using Euclidean links for a given landscape. Then for each of the eight cost surfaces that corresponded to that landscape, I computed the least-cost links between all pairs of nodes that were present in the Euclidean minimum planar graph. This produced an additional ten graphs that all had identical node and link sets; however, the length, the effective length, and the spatial location of the links could differ depending on the cost surface. In total, for each of the 2 000 landscapes and 8 sets of cost values, I produced 18 000 graphs (2 000 using Euclidean links and 16 000 using least-cost links). Results were analyzed using

ANOVA (e.g., Li and Reynolds, 1994; Matter, 2003; Hufkens et al., 2008) in order to Chapter 4. Sensitivity of least-cost habitat graphs 79 examine main effects, two-way interactions, and higher-order interactions among the four factors. No transformations of the data were required in order to meet the normal- ity assumptions of ANOVA.

4.3.4 Measuring the sensitivity of graphs with least-cost links

For each graph produced with least-cost links, I calculated a standardized measure of the spatial deviation of least-cost links attributable to the corresponding cost surface.

This standardized measure, referred to as the total spatial deviation of links, was cal- culated as the difference between the geographic length of the least-cost link and the

Euclidean link for each pair of connected patches, summed over all pairs of connected patches in the landscape graph. Large values indicate the presence of sinuous least-cost links that deviate from straight-line Euclidean links, whereas small values indicate that least-cost links closely follow the straight-line Euclidean links. The sinuosity of least- cost links will depend on both the spatial pattern of the habitat and matrix landcover types as well as the relative cost values assigned to each. This is why I ran a factorial experiment with three factors controlling landscape spatial pattern (degree of fragmen- tation of primary habitat, amount of hospitable matrix, and degree of fragmentation of hospitable matrix) and one factor controlling cost values (set of relative cost values) to test which factors are most important in determining total spatial deviation of least- cost links (Table 4.2).

4.4 Results

The results of the 4-factor ANOVA (Table 4.4) showed a significant three-way interac- tion between matrix composition (HM COV), habitat fragmentation (H FRAG), and the set of relative cost values assigned to different landcover types (COST). This signif- icant interaction precluded any further significance testing for the two-way interactions Chapter 4. Sensitivity of least-cost habitat graphs 80

or main effects; however, I was able to observe that the ANOVA clearly did not indi-

cate significance for any two-way interactions including HM FRAG or its main effect

(Table 4.4).

The three-way interaction can be examined visually by plotting the interaction be-

tween two of the factors at each level of the third factor (Figs. 4.3 and 4.4). In all of

the two-way interaction plots the most striking result was the consistent differentiation

between sets of cost values with odd numbers (C1, C3, C5, C7) versus even numbers

(C2, C4, C6, C8). Cost sets with odd numbers had small differences between the rel-

ative costs of hospitable and inhospitable matrix landcover types, whereas even num-

bered cost sets had large differences. This result is intuitive because larger differences

between matrix landcover cost values would expectedly cause least-cost links to deviate

more strongly from straight-line links. The difference in cost values between habitat

and the matrix was evidently not important despite relative values that differed 100

fold. The interaction line for C2 lies between the odd- and even-numbered cost sets

and shows dampened peaks consistent with the odd-numbered cost sets.

The two-way interaction between HM COV and COST displayed a similar pattern for both low and high levels of H FRAG (Fig. 4.3); however, spatial deviation values

were five times larger where H FRAG was high (Fig. 4.3b) compared to where it was

low (Fig. 4.3a). Cost-value sets C1, C3, C5, C7 and sets C2, C4, C6, C8 differed in

their response to increasing HM COV for both levels of H FRAG. Two thresholds of

HM COV are apparent in these interaction plots. First, there is an abrupt change in

the behavior of C2, C4, C6, and C8 when HM COV equals 30%. Below this level of

HM COV the spatial deviation of links increases with increasing HM COV and above

this level of HM COV the spatial deviation of links decreases or stays the same with

increasing HM COV. Second, when HM COV is above 50% all cost-value sets become consistent in their response to increasing HM COV.

The two-way interaction between H FRAG and COST also showed a difference be- Chapter 4. Sensitivity of least-cost habitat graphs 81

Table 4.4: Effects of matrix composition (HM COV), habitat fragmentation

(H FRAG), matrix fragmentation (HM FRAG), and the relative cost values (COST) on the mean spatial deviation of least-cost links. Presented are F-ratios and P-values

from the four-factor ANOVA.

Source df F-ratio P-value

HM COV 4 12821.583 <0.001

H FRAG 1 46350.009 <0.001

HM FRAG 1 1.145 0.285

COST 7 14319.069 <0.001

HM COV*H FRAG 4 5616.703 <0.001

HM COV*HM FRAG 4 0.006 0.999

H FRAG*HM FRAG 1 1.265 × 10−05 0.997

HM COV*COST 28 2510.185 <0.001

H FRAG*COST 7 6923.999 <0.001

HM FRAG*COST 7 0.004 1.000

HM COV*H FRAG*HM FRAG 4 9.306 × 10−06 1.000

HM COV*H FRAG*COST 28 1135.699 <0.001

HM COV*HM FRAG*COST 28 2.652 × 10−05 1.000

H FRAG*HM FRAG*COST 7 2.033 × 10−05 1.000

HM COV*H FRAG*HM FRAG*COST 28 4.076 × 10−05 1.000

Residuals 15840 Chapter 4. Sensitivity of least-cost habitat graphs 82

Figure 4.3: Interaction plots for the percentage of hospitable matrix (HM COV) versus the set of relative cost values (C1 - C8) when habitat fragmentation is either a) low

(H FRAG = 0.05) or b) high (H FRAG = 0.5). Mean values of the total spatial devia- tion of least-cost links is plotted for each of the cost sets at each level of HM COV. Chapter 4. Sensitivity of least-cost habitat graphs 83

Figure 4.4: Interaction plots for the fragmentation of habitat (H FRAG) versus the

set of relative cost values (C1 - C8) plotted separately at each percentage of hospitable

matrix (HM COV; a - e). Mean values of the total spatial deviation of least-cost links is plotted for each of the cost sets at each level of H FRAG. Chapter 4. Sensitivity of least-cost habitat graphs 84

tween the responses of cost-value sets C1, C3, C5, C7 and sets C2, C4, C6, C8 for all

values of HM COV (Fig. 4.4). As H FRAG increased, the spatial deviation of least- cost links also increased; however, this increase was very slight for cost-value sets C1,

C3, C5, C7 and was dramatic for cost-value sets C2, C4, C6, C8. Differences among cost sets were in general much larger when H FRAG was high. The rate at which the

spatial deviation of links in C2, C4, C6, and C8 increased in response to H FRAG dif-

fered depending on the level of HM COV. When HM COV was 10%, their slopes were

visibly not parallel, indicating that they had different responses to increasing H FRAG.

When HM COV was higher than 10% their slopes were generally parallel with the ex-

ception of C2.

4.5 Discussion

This study confirmed that connectivity assessments based on least-cost routes do reveal

effects of landscapes on potential movement pathways of organisms. Of interest, how-

ever, was the question of how sensitive these connectivity assessments were to the pa-

rameterization of the cost surface. Only a handful of connectivity studies that assume

least-cost movement behavior have performed sensitivity analyses on the landcover cost

parameter and most of these sensitivity analyses are performed on a single landscape

(Table 4.1). Although these studies employ a variety of connectivity measures and test

for their sensitivity with varying degrees of rigor, the majority did reveal some effects

of the cost values on their results. My sensitivity analysis is not comprehensive but it

is the most rigorous of its kind in terms of examining the interaction effects between

landscape structure and cost-surface parameterization on least-cost route analyses. I

found that indeed there were significant interactions; hence, the sensitivity of least-cost

links to relative cost values changed depending on the level of habitat fragmentation

and the amount of hospitable matrix. Gonzales and Gergel (2007) also found that the Chapter 4. Sensitivity of least-cost habitat graphs 85 results of their least-cost route analyses differed in fragmented and continuous land- scapes.

My analyses clearly showed that the least-cost links were more sensitive to the difference in cost values between hospitable and inhospitable matrix landcover types than between habitat and hospitable matrix landcover. I restricted my analyses to the extreme differences in relative cost values between matrix landcover types that had been reported in the literature (1.5 and 100 fold; Table 4.1) so that I could look at interactions among landscape structure and cost sensitivity. I acknowledge that more work needs to be undertaken to define a potential range of cost values that could pro- duce consistent results in least-cost connectivity assessments in a variety of landscapes.

Schadt et al. (2002) confirmed that their ability to classify connections among habi- tat patches in terms of exchange of animals and mortality risk was only sensitive to cost values when habitats within the matrix had “very low” values. In their study, they varied the cost of a hospitable matrix landcover type with relative cost values of 4, 7, 10, 20, 30, 100, and 500 whereas inhospitable matrix was assigned a constant cost of 1000. Hospitable matrix cost values from 7 to 500 produced similar results and were deemed biologically plausible, but the smallest cost value of 4 changed their re- sults and was dismissed as being below the biologically realistic range. Chardon et al.

(2003) and Verbeylen et al. (2003) also found that models assuming least-cost move- ment fitted patch-occupancy data better than the models assuming Euclidean move- ment for a wide range of cost values. From this result, they concluded that all chosen parameters and ratios were in the range appropriate for predicting the occurrence of their study species. However, they also found that the predictive power of their mod- els were sensitive to the exact quantification of the cost surface. This supports the idea that a biologically-realistic range of cost values may produce satisfactorily consistent results for coarse measures of connectivity, although some sets of cost values will im- prove accuracy of precise predictions. In light of these findings and given the extreme Chapter 4. Sensitivity of least-cost habitat graphs 86 cost values assigned to matrix landcover types in my study, I expected to observe some differences in the spatial locations of least-cost links. Indeed, least-cost links did fol- low larger spatial detours from straight-line links where the difference in cost values between matrix cover types was large (e.g., Fig. 4.2). The weaker effect of the differ- ence between cost values assigned to habitat and hospitable matrix was unexpected as it has not been explicitly tested in previous least-cost route sensitivity studies. This re- sult is not entirely surprising however, given that my sensitivity analysis of least-cost routes was combined with a graph-analysis approach. In the graph approach least-cost links did not cross nodes; hence, I would expect that multiplying all cost values ex- cept for the cost of habitat (which is fixed at 1) by a positive integer would result in the same spatial pattern of least-cost links, but the effective distance of links would be correspondingly larger. Therefore, I reiterate that the most important relative cost val- ues are those between landcover types in the matrix when evaluating the sensitivity of least-cost routes implemented within a habitat graph modelling approach.

My results also showed that the sensitivity of least-cost links to relative cost val- ues was modified by the composition of the matrix. The spatial deviation of least-cost links from straight-line links peaked where the percentage of hospitable matrix in the landscape was 30%. For higher percentages of hospitable matrix, different sets of rel- ative cost values produced more consistent least-cost links. Conversely, the degree of fragmentation of the matrix landcover types did not have an effect on the sensitivity of least-cost links to relative cost values. This result is consistent with results from empir- ical studies demonstrating that the effects of fragmentation are generally much weaker than the effects of habitat amount on various measures of species movement, distribu- tion, and persistence (Fahrig, 2003, and references therein). I set the amount of habi- tat in all of my simulated landscapes as 12% to reflect common conservation targets

(Pressey et al., 2003) and found that for this proportion of habitat, the degree of frag- mentation of the habitat did indeed affect the sensitivity of least-cost routes. In this Chapter 4. Sensitivity of least-cost habitat graphs 87

respect, the least-cost habitat graph model is behaving consistently with the observa-

tion that the effects of fragmentation increase with decreasing habitat amount in the

landscape (Andr´en,1994, 1999) because I have a low proportion of habitat in my land-

scapes. Indeed, the quality of landcover types has also been shown to modify the rela-

tionship between the size of habitat patches and population abundances (M˝onkk˝onen

and Reunanen, 1999). The complex interaction among habitat fragmentation, amount

of hospitable matrix, and relative cost values in my study further emphasize the impor-

tance of quantifying all three (pattern, amount, and quality) when making assessments

of potential movement through heterogeneous landscapes.

Although a wide variety of values have been used to quantify habitat cost values

for many different species in the connectivity literature (see Table 4.1), very few stud-

ies have empirically estimated relative cost values. There have been two studies, how-

ever, which have addressed this question with very different approaches (Ricketts, 2001;

Stevens et al., 2004). Stevens et al. (2004) conducted a manipulative, laboratory-based

experiment to compare physical movement ability through 5 landcover types corre-

sponding to sand, cement, field grass, and forest for the toad Bufo calamita. They es- timated cost values of 1, 1.5, 2, and 2.5 for these habitats, but these cost values only measure a single aspect of fitness, which limits their direct applicability to connectiv- ity analyses. Ricketts (2001) used field surveys to estimate the difference between wil- low and conifer costs for six butterfly taxa. He found that the willow landcover had a cost value of 1.7, 0.9, 1.4, 2.0, and 1.1 for five of the six taxa studied relative to the meadow cost of 1. For four taxa, conifer cost was between 3 and 12 times higher than the willow cost. The small differences in landcover cost estimates from these studies raise questions about the large relative cost values that are generally assigned by ex- pert opinion (see Table 4.1); however, it is important to note that the studies were all examining different species with different ecological sensitivities which could account for the drastically different cost values. Manipulative methods for the empirical estima- Chapter 4. Sensitivity of least-cost habitat graphs 88 tion of landcover cost have also been suggested by B´elisle(2005), such as translocation and playback experiments, food titration experiments, and manipulating feeding and breeding-site locations. These types of empirical efforts coupled with my work on the sensitivity of results to the cost values will be essential for the wide-spread acceptance of graph-theoretic connectivity analyses that assume least-cost movement behavior as a tool to inform spatial planning decisions for conservation.

4.6 Conclusion

Despite their sensitivity to the parameterization of the cost surface, least-cost routes are increasingly being coupled with graph-theoretic techniques to assess connectiv- ity in a diversity of applications. Easily accessible GIS software that allows least-cost routes to be computed rapidly has led to a proliferation in their use without examining key assumptions that may affect the ecological relevance of the conclusions based on their results (Gonzales and Gergel, 2007). In my systematic study I considered how the interactions between quality, amount, and fragmentation of habitat and matrix-cover types influence connectivity assessments that are based on least-cost routes and graph- theoretic analyses. I found that landscape structure affects how sensitive these connec- tivity assessments are to the quality (relative cost values) of landcover types, with the largest sensitivity occurring in fragmented landscapes. To overcome this sensitivity, I suggest identifying multiple low-cost paths between pairs of habitat patches that collec- tively delineate spatial zones (areas) accessible for probable movement within the inter- vening landscape (e.g., Sutherland et al., 2007). Pinto and Keitt (2009) have recently presented two methods for identifying multiple paths with similar costs, the Condi- tional Minimum Transit Cost method (see also Walker and Craighead, 1997; Halpin and Bunn, 2000; Theobald, 2006) and the Multiple Shortest Path method. These spa- tial movement zones may be more robust to variation in cost values because the uncer- Chapter 4. Sensitivity of least-cost habitat graphs 89 tainty in least-cost movement is incorporated into the delineation of the zones (Magle et al., 2009; Pinto and Keitt, 2009). Additional spatial properties of these movement zones such as their geometry and area can provide additional information about the quality of connections between habitat patches (Theobald, 2006; McRae et al., 2008).

Coupled with graph-theoretic analyses, these multiple-link movement zones will be par- ticularly useful for guiding conservation planning decisions that must often be made quickly and with sparse data on disperser behavior. Chapter 5

Connectivity for conservation: A framework to classify habitat-network-connectivity statistics

5.1 Abstract

Graph theory, network theory, and circuit theory are increasingly being used to quan- tify multiple aspects of connectivity in habitat networks. There has been a prolifera- tion of connectivity statistics for habitat networks resulting in over sixty statistics for ecologists to now choose from. Conceptual clarification on the ecological meaning of these network connectivity statistics and their interrelationships is overdue. I present a classification framework that categorizes connectivity statistics based on the com- ponent of connectivity that they quantify (i.e., dispersal likelihood, route redundancy, route vulnerability, and habitat area connected) and the level of the habitat network to which they apply. The value of this framework lies in its ability to enable more

90 Chapter 5. Classification of network-connectivity statistics 91 informed choices and applications of network connectivity statistics. This will allow graph, network, and circuit analyses to vastly improve our ability to design and man- age connected landscapes.

5.2 Introduction

The ability of individuals to move among habitat patches and populations is being un- dermined by ongoing habitat fragmentation, land-use intensification, and biotic homog- enization (Saunders et al., 1991; Fahrig, 2003; Olden and Poff, 2003; Fischer and Lin- denmayer, 2007). Reduced and impeded movements have significant consequences for biodiversity conservation (Taylor et al., 1993; Kareiva and Wennergren, 1995; Moila- nen et al., 2005; Crooks and Sanjayan, 2006; Damschen et al., 2006; Minor and Ur- ban, 2008). Maintaining movements of individuals in the short- to medium-terms en- ables juvenile dispersal, recolonization of empty habitat patches, seasonal migration, and metapopulation persistence (Gilpin and Soul´e,1986; Hanski, 1998a). Maintaining movements of individuals in the long-term enables range shifts in response to climate change and conserves genetic diversity required for evolutionary adaptation (Schemske et al., 1994; Minor and Urban, 2008). Quantifying the degree to which the landscape promotes or hinders movements among patches of habitat for a given species, hereafter habitat connectivity (Fischer and Lindenmayer, 2007), is therefore essential to inform conservation-management decisions (Calabrese and Fagan, 2004).

The graph-theoretic approach is a method for estimating habitat connectivity that is rapidly increasing in popularity in the disciplines of landscape ecology and conser- vation biology (Fig. 5.1 Urban and Keitt, 2001; Fall et al., 2007). This popularity can be attributed to three primary strengths of graph theory: (1) its efficiency in charac- terizing connectivity at broad spatial scales in landscapes with many habitat patches

(Urban et al. 2009); (2) its ability to balance data requirements with information con- Chapter 5. Classification of network-connectivity statistics 92 tent (Calabrese and Fagan, 2004); and (3) its flexibility to incorporate additional in- formation about relevant aspects of a species’ biology into connectivity assessments

(Fall et al., 2007; Minor and Urban, 2008). The graph-theoretic approach represents the connectivity of a set of habitat patches as a “habitat graph” - a collection of nodes

(habitat patches) and links that connect pairs of nodes (representing the potential or frequency of movement between habitat patches; Box 5.1). The way in which nodes and links are defined will determine whether the habitat graph represents structural, potential, or functional connectivity among habitat patches (Calabrese and Fagan,

2004; B´elisle,2005; Fagan and Calabrese, 2006). Representing structural connectivity requires, for example, that links simply encode information about the physical adja- cency of habitat patches or the physical distances among them. Additional informa- tion about the focal species’ dispersal abilities, such as its maximum dispersal-distance threshold, may be used to eliminate links that exceed that threshold distance and pro- duce a representation of potential connectivity for the focal species (e.g., Brooks et al.,

2008; Pascual-Hortal and Saura, 2008). Finally, if data are available on actual move- ment patterns of individuals among habitat patches, then links can be defined based on these observed movements and the corresponding habitat graph would represent functional connectivity. In the absence of observed movement data, only graphs repre- senting landscape heterogeneity can provide some insights into potential connectivity.

However, not all species or individuals respond similarly to different landscape con-

figuration and composition. It is therefore important to evaluate habitat connectivity based on different properties of the graph.

Over the past decade there has been extensive proliferation of connectivity statis- tics for habitat graphs emanating from the related fields of graph, network, and circuit theory (over 60 statistics; see Tables 5.2 and 5.3). These statistics were either devel- oped specifically for the application to habitat-connectivity assessments (Urban and

Keitt, 2001; Pascual-Hortal and Saura, 2006; Saura and Pascual-Hortal, 2007, e.g.,) Chapter 5. Classification of network-connectivity statistics 93 or were adapted from other fields such as social sciences, transportation theory, com- munication theory, and epidemiology (e.g., Fortuna et al., 2006; Brooks et al., 2008).

In the absence of any selection guidelines, this proliferation of connectivity statistics poses a challenge for ecologists trying to select one or a few statistics for their graph- based connectivity assessments. The result is that most studies use different connectiv- ity statistics making the results difficult to interpret and compare. A framework that classifies graph-based connectivity statistics is therefore necessary so that informed de- cisions can be made regarding the choice of statistics for any particular assessment of habitat connectivity.

My goal is to devise a classification framework of connectivity statistics for graph, network, and circuit analyses of habitat connectivity. I begin by reviewing the meth- ods of graph, network, and circuit theories with an emphasis on the ways in which they incorporate spatial information about the habitat patches and the surrounding matrix in the definition of nodes and links. I then present a classification framework that cate- gorizes connectivity statistics based on the level of the graph to which they apply (i.e., element, neighbourhood, cluster, and network levels) and the component of connectiv- ity that they quantify (i.e., dispersal likelihood, route redundancy, route vulnerability, and habitat area connected). I conclude with a proposed extension of these static habi- tat graphs that explicitly incorporates temporal changes in habitat connectivity (i.e., spatio-temporal connectivity).

5.3 Background

“Habitat networks” (Keitt, 2003; Fortuna et al., 2006; Theobald, 2006; Brooks et al.,

2008; Economo and Keitt, 2008) and “habitat circuits” (McRae, 2006; McRae and

Beier, 2007; McRae et al., 2008) are special forms of habitat graphs with distinct prop- erties that make them especially well-suited to represent potential and functional habi- Chapter 5. Classification of network-connectivity statistics 94

Figure 5.1: Number of published studies that use graph, network, or circuit theory to quantify habitat connectivity. The studies are ordered by their year of publication and include five review articles, three conference proceedings, and a user manual. Studies have been categorized based on their self-ascribed method of analysis: graph theory, network theory, or circuit theory. Note that these self-ascribed labels have been incon- sistently applied and do not necessarily correspond to my definitions (Box 5.1). Chapter 5. Classification of network-connectivity statistics 95 tat connectivity (Box 5.1). A “habitat network” is defined as a habitat graph with nodes and links that are assigned positive numeric weights that reflect their influence on movements (i.e., nodes could be weighted by their expected or observed number of emigrants and links could be weighted by their expected or observed number of mi- grants that successfully travel their length; Theobald, 2006; Brooks et al., 2008). A

“habitat circuit” is defined as a habitat network with nodes that are connected by spe- cial links made up of multiple, alternative movement routes (McRae, 2006; McRae and

Beier, 2007; McRae et al., 2008). Hereafter, I use the term habitat network as an intu- itive, general reference to a graph-based model of the potential or functional connectiv- ity among habitat patches. I make specific reference to habitat graphs, habitat circuits, and their respective analytic methods as necessary (Box 5.1).

Once the nodes and links of a habitat network have been defined, network connec- tivity statistics can be calculated to quantify patterns of compartmentalization and connectivity within the network or across the entire network (Table 5.1; Keitt et al.,

1997).

Box 5.1. Graphs, networks, and circuits Habitat connectivity can be represented by graphs, networks, and circuits that differ with respect to the ecological information they encapsulate. These different representations of the habitat system will be suited to different applications de- pending on the data available and the questions of interest. The distinction be- tween graphs, networks, and circuits is not well defined and the terms are often used interchangeably in different disciplines. Some distinctions can be drawn how- ever based on either the models themselves or on the methods of analysis. In general, networks and circuits can be thought of as special types of graph models with restricted definitions and applications. In its most basic form, a graph is a set of nodes connected to some degree by links that join pairs of nodes (Harary, 1969). The term “network” is used synonymously with the term “graph” (Borgatti et al., 2009) but it should in fact refer exclusively to weighted graphs (Gross and Yellen, 2004; Diestel, 2006, Box2;), in which nodes and links have posi- tive numeric weights. Nodes and links may be assigned weights based on attributes that affect connectivity, such as the size of nodes and the length of links in the case of habitat networks. In general, networks are used in an applied context and repre- sent real-world systems (Newman et al., 2006). A circuit (specifically an electrical Chapter 5. Classification of network-connectivity statistics 96

circuit) is a network or multigraph (Box 5.2) in which nodes are connected by spe- cial links made up of one or more resistors (electrical components that conduct cur- rent; McRae et al., 2008). Such parallel connections may correspond to alternative movement routes between pairs of connected nodes in a habitat networks. Just as networks and circuits are graphs with specialized properties or appli- cations, network theory and circuit theory are both derived from the mathemati- cal subdiscipline of graph theory which is broadly defined as the study of graphs (Harary, 1969). Graph theory has been the primary mathematical language used to describe the structural properties and behavioral characteristics of networks to date (Newman et al., 2006); however, its scope extends beyond applied networks and also includes a large body of work in algebraic analyses, geometric problems, and pure mathematics (Urban et al., 2009). Network theory began as an applica- tion of graph theory carried out in a number of applied research areas including so- cial networks, electrical circuits, transportation systems, communication networks, epidemiology, and now ecological habitat networks (Brandes and Erlebach, 2005). One of the first applications of network theory was to model ecological food webs and, inturn, this application has greatly contributed to the development of net- work theory (Bascompte, 2007). It has recently set itself apart from graph theory by establishing a well-defined research agenda focused on properties of real-world networks, their structural dynamics, and the relationship between their structure and function (Newman et al., 2006). Hence, network theory provides a mathemat- ical foundation for studying the effects of habitat network structure on dispersal processes and ultimately biodiversity persistence. Circuit theory (specifically elec- trical circuit theory) is a specific application of network theory dedicated to quan- tifying voltage, current, conductance, and resistance in circuits (Dorf and Svoboda, 2003). Circuit analyses are particularly useful for measuring habitat connectivity for species with dispersal behaviors that respond positively to increasing connec- tions and redundancy. Measures of connectivity in circuit analyses incorporate all possible pathways and the cumulative flow between nodes (McRae, 2006; McRae and Beier, 2007; McRae et al., 2008).

Box 1 Figure 1. Representations of habitat connectivity that differ with respect to the amount of ecological information that they incorporate. Habitat patches (black polygons) are connected by links (black lines) that cross hospitable (grey) and inhospitable (white) matrix cover types. a) A habitat graph connects patch centroids without incorporating a lot of spatial and ecological information about Chapter 5. Classification of network-connectivity statistics 97

nodes and links. b) A habitat network connects patch edges by least-cost links that incorporate information about matrix heterogeneity. Additional node and link at- tributes may also be included by assigning weights. c) A habitat circuit connects patches with multiple links thereby incorporating additional spatial information about the matrix. .

5.4 Development of methods to construct habitat

networks

Network theory belongs to the branch of mathematics that studies connections among discrete objects known as graph theory (see Box 5.1). Developments in the application of network theory to quantify habitat connectivity have been made primarily using the language of graph theory (Urban and Keitt, 2001; Fall et al., 2007; Pinto and Keitt,

2009). Initially, graph-theoretic approaches were proposed to quantify habitat connec- tivity based on very simple habitat graph models (Marcot and Chinn, 1982; Cantwell and Forman, 1993; Forman, 1995; Keitt et al., 1997; Marcot, 1998). These simple habi- tat graphs were non-directed, unweighted graphs (Box 5.2), in which habitat patches were represented as dimensionless nodes and links were represented as lines connect- ing nodes. Any information about the spatial location, length, shape, and quality of habitat patches and of links was disregarded. Therefore, estimates of habitat connec- tivity based on these simple habitat graphs considered only the presence or absence of connections between habitat patches (topology; Box 5.2) instead of the strength of connections.

Further developments of the network-theoretic application have resulted from the inclusion of more information about habitat suitability, landscape permeability, and movement behavior into habitat graphs (van Langevelde et al., 1998). The most straight- forward means of including additional information into simple habitat graph models is Chapter 5. Classification of network-connectivity statistics 98 to assign positive numeric weights to the nodes and links. Nodes can be weighted to reflect habitat-patch properties that may influence immigration and emigration such as patch area (e.g., Keitt et al., 1997; Bunn et al., 2000; Estrada-Pena, 2002; Miller and

Russell, 2004) or patch quality (e.g., Jord´an,2003; Minor and Urban, 2007; Saura and

Pascual-Hortal, 2007). Links can be weighted to reflect link properties that may af- fect dispersal such as the geometric length of the link or the effective length of the link based on estimated movement costs of underlying landcover types (e.g., Halpin and

Bunn, 2000; Jord´an,2000; Urban and Keitt, 2001; Keitt, 2003; Rothley, 2005; O’Brien et al., 2006; Fall et al., 2007). Link weights have also been used to describe the dis- persal probability between two nodes by assuming a negative exponential relationship between dispersal probability and link length (e.g., Urban and Keitt, 2001; Keitt, 2003;

Brooks et al., 2008; Treml et al., 2008). These weighted habitat graphs have been de- scribed using the language of both graph theory (Bunn et al., 2000; Urban and Keitt,

2001; Brooks, 2006, e.g.,) and network theory (Keitt, 2003; Fortuna et al., 2006; Bodin and Norberg, 2007; Brooks et al., 2008, e.g.,). More sophisticated means of incorpo- rating additional information into habitat networks focus on maintaining spatial ref- erencing of nodes and links during the construction of habitat networks (Theobald,

2006; Fall et al., 2007). These more sophisticated methods explicitly account for habi- tat patch shape and landscape structure in two ways: (1) nodes are treated as two- dimensional patches with fixed spatial locations; and (2) links between nodes con- nect patch perimeter to patch perimeter and follow geo-referenced least-cost routes.

Connectivity of these habitat networks has been measured in terms of the probabil- ity of movement among habitat patches and the area of habitat connected (Saura and

Pascual-Hortal, 2007, and references therein).

The most recent development of the network-theoretic approach has been to treat links as multiple routes instead of single, least-cost routes (McRae, 2006; Theobald,

2006; McRae and Beier, 2007; McRae et al., 2008; Pinto and Keitt, 2009, reviewed in Chapter 5. Classification of network-connectivity statistics 99

Urban et al., 2009). Two methods have been proposed for identifying multiple routes between habitat nodes: the conditional-minimum-transit-cost method (CMTC method;

Theobald, 2006; Pinto and Keitt, 2009) and the multiple-shortest-paths method (MSP method; Pinto and Keitt, 2009). Both of these methods identify a large set of routes between a pair of habitat patches but only retain the subset that are within a given distance of the length of the least-cost route. The CMTC method identifies a set of routes such that each path minimizes the CMTC through a different location on the landscape. The CMTC is measured as the least-cost distance between two habitat nodes contingent upon the least-cost path crossing a given location on the landscape

(i.e. a given grid cell in a raster map). The MSP generates a set of paths between two nodes by repeated permutation of the least-cost route. Route segments are randomly deleted and the least-cost route is re-calculated in an iterative fashion. These meth- ods allow us to create multi-route habitat networks (or multigraphs or circuits; Box

5.2) which provide important spatial information about movement options through the matrix and allow for increased flexibility in connectivity conservation planning (Urban et al., 2009).

Box 5.2. Definitions of Network Terminology Adjacency matrix - a binary matrix in which each entry is defined aij = 1 if nodes i and j are share a link, otherwise aij = 0. The matrix has one row and one column for each node. Circuit (electrical) - a network or multigraph in which nodes are connected by special links made up of one or more resistors (electrical components that conduct current). Circuit theory (electrical) - a specific application of network theory dedicated to quantifying voltage, current, conductance, and resistance in electrical circuits. Cluster/component - a group of interconnected nodes for which a path exists between every pair of nodes. Cost value - a value assigned to each habitat type in a landscape that reflects the quality of the habitat in terms of ecological costs incurred by an individual moving through that habitat type. The cost value of each habitat type is normalized by the cost value of the preferred habitat type. Degree of a node - the number of links incident to that node. Chapter 5. Classification of network-connectivity statistics 100

Degree distribution - the frequency distribution of node degrees for all nodes in a network. Directed graph/digraph - connected nodes in a directed graph have two links, one in each direction. Digraphs represent directional asymmetries. Dispersal probability matrix - consists of the probabilities that an individual disperser will move between pairs of nodes, for all node pairs in the network. Distance matrix - consists of the metric distances (either geographic or cost) be- tween pairs of nodes, for all node pairs in the network. Geodesic distance - the length of the geodesic path measured as either the topo- logical or the metric distance. Geodesic path - the shortest path between two nodes calculated as either 1) the path that traverses the shortest number of links between two nodes or 2) the path that minimizes the sum of the link weights of the links in the path. Graph - a set of nodes connected to some degree by links that join pairs of nodes. Graph theory - the study of properties of graphs. Incidence matrix - a binary matrix in which each entry is defined as cij = 1 if node i is incident upon link j, 0 otherwise. The matrix has one row for each node and one column for each link. Least-cost surface - a map of the cost values associated with each habitat type in a landscape. Least-cost route - a link or path that minimizes the cumulative cost values along its length. Link/edge - a network element that connects nodes. Metric distance - distance between two nodes in a network can be measured in geometric distance units or cost distance units based on the estimated movement costs of landcover types that are traversed. Motif - a small repeating pattern (sub-graph) within a network. Multigraph - a graph which has multiple or parallel links between a pair of nodes. Neighborhood - the set of nodes that are within a given number of connections from a focal node (primary neighborhood corresponds to nodes within one connec- tion, secondary neighborhood corresponds to nodes within two connections, etc.). Network - a weighted graph or digraph that carries all the information about the inter-connection of the elements in a real-world system. Network theory - the study of properties of networks with an emphasis on their structural and behavioral dynamics. Network static - single numbers or series of numbers that capture the relevant characteristics of a network. Node/vertex - distinct elements within a network that are connected by links if they interact in some way. Path - a route that indirectly connects nodes along a series of connected links and nodes in which no node is visited more than once. Planar graph or network - a graph or network is planar if it can be drawn on a plane so that the links only intersect at the nodes (i.e., they do not cross between Chapter 5. Classification of network-connectivity statistics 101

nodes). Redundancy matrix - consists of the number of alternative routes that exist be- tween pairs of nodes, for all node pairs in the network. Topological distance - distance between two nodes in a network is the number of links between these nodes. Topology - the configuration of nodes and links in a network. Undirected graph - connected nodes in an undirected graph have a single link representing symmetric connectivity in both directions. Produces a symmetric ad- jacency matrix. Unweighted graph - a graph in which links are binary: assigned a value of one if the link exists and zero otherwise. Vulnerability matrix - consists of the number of cut-nodes, cut-links, or cut-grid cells that exist along routes between pairs of nodes, for all node pairs in the net- work. Weighted graph - nodes and links are assigned positive numeric integers describ- ing their contribution to connectivity of the graph. Sometimes used synonymously with the term “network”. .

5.5 Quantifying connectivity in habitat networks

Due to the large size and complexity of habitat networks, it is necessary to condense and summarize the relevant information via the use of graph and network statistics

(Box 5.2). A network statistic is either a single number or a series of numbers that quantifies some property of a network (Brinkmeier and Schank, 2005). Through the years, a wide variety of network statistics have been proposed to quantify the connec- tivity of habitat networks. I reviewed the literature covered by the ISI Web of Knowl- edge Science Citation Index (www.isiknowledge.com) database using the following word combinations in a topic search: graph, network, or circuit in combination with habitat connectivity or landscape connectivity. I found 51 relevant articles dating from January

1982 to May 2009 that have cumulatively proposed 60 network connectivity statistics for quantifying habitat connectivity. These network connectivity statistics have been assembled in Tables 5.1 and 5.2.

Many of these statistics have been imported or adapted from other disciplines. For Chapter 5. Classification of network-connectivity statistics 102 example, the Gamma Index (i.e., the number of links in a network divided by the num- ber of links in the corresponding planar network; Table 5.2) was first imported from transportation geography by Ricotta (2000) to quantify the connectivity of different vegetation types in a landscape and later modified by Acosta et al. (2003) to include qualitative differences in the conservation values of vegetation types. Other statistics have been developed specifically for ecological applications such as the Integral Index of Connectivity (Table 5.3; Pascual-Hortal and Saura, 2006) which was proposed as a means of integrating habitat area and habitat connectivity in a single measure.

Network statistics that quantify habitat connectivity can broadly be divided into topological indices (Table 5.2) and weighted indices (Table 5.3). Topological network statistics only consider the presence or absence of a link between nodes during their calculation. They characterize the basic structure of a network based on the qualitative pattern of connections among nodes (Table 5.2). Weighted network statistics consider the variation and strength of connections between nodes by including node and link weights during their calculation. They characterize connectivity patterns in the net- work based on ecological differences between nodes and links (Table 5.3).

5.6 Classification framework of habitat network con-

nectivity statistics

To organize this large number of network-connectivity statistics, I have classified them based on two criteria: (1) the network level at which they can be applied (i.e., element, neighborhood, cluster, and network; Fig. 5.2); and (2) the specific components of con- nectivity that they measure (Table 5.1). To facilitate this classification, the compo- nents of connectivity are simplified into four primary categories: inter-patch dispersal likelihood, route redundancy, route vulnerability, and habitat availability. The pur- pose of this classification scheme is to illustrate the parallels that exist between com- Chapter 5. Classification of network-connectivity statistics 103 ponents of habitat connectivity that can be measured at different levels of the habi- tat network. This framework will allow researchers and practitioners to quantify habi- tat connectivity at multiple levels of analysis and make a more informed selection of habitat-network-connectivity statistics based on the component of connectivity that they measure.

5.6.1 Network levels of analysis

Distinct levels of analysis exist within networks (Fig. 5.2; Wasserman and Faust, 1994;

Brandes and Erlebach, 2005) and network connectivity statistics have been developed for each of the distinct levels. There are connectivity measures, for instance, pertain- ing exclusively to the elements of the network themselves (i.e., nodes and links). At this “element level”, connectivity statistics focus on properties such as the number of links per node (i.e., node degree) or the importance of a node based on its structural position within the network (i.e., node centrality). “Neighborhood-level” statistics as- sess the connectivity among a set of nodes that are within a given number of connec- tions from a focal node (neighborhood; Box 5.2). For example, a neighborhood analy- sis could examine the relationship between the degree of a node and the degrees of its neighbors (i.e., connectivity correlation). “Cluster-level” statistics examine groups of interconnected nodes (cluster; Box 5.2), focusing perhaps on the number of nodes con- nected (i.e., cluster order) or the average length of the shortest-path connecting node pairs (i.e., characteristic path length). Finally, there are “network-level” connectivity measures that focus on the patterns of connectivity among all nodes and links in the network (e.g., the mean node degree or the number of clusters).

Higher levels of analysis, such as the cluster and network levels, often summarize the distribution of connectivity properties at lower levels with frequency plots or mea- sures of central tendency and spread. These high-level network descriptors can be ei- ther single-valued statistics or distributions (Brinkmeier and Schank, 2005). Basic Chapter 5. Classification of network-connectivity statistics 104 Assymetry Connectance CV of cluster order Node degree distribution Diameter of largest cluster Expected dispersal flux Gamma index Mean cluster size Mean node degree Number of clusters Order Order of largest cluster Order of smallest cluster Recruitment Size Traversability Total number of cutTotal nodes number of cut links Area index Class coincidence probability Correlation length Expected cluster size (area) Integral index of connectivity Landscape coincidence probability Max. connected local pop.Probability of size connectivity Average path strength Characteristic path length Cluster order Cluster size Diameter Harary index Path strength Wiener index Number of cut nodes Number of cut links Cluster area Level of Network Analysis Clustering coefficientEcologically scaled connectivity Meshedness Betweeness centrality Closeness centrality Dispersal flux Eccentricity Node degree Node degree correlation Node depth Node in degree Node out degree Node influx Node outflux Reachability index Commute time Effective resistance Network flow Route (link) redundancy Current densityNode area Modified incidence Connectivity correlationfunction Reliability Quality-weighted area Element Neighbourhood Cluster Network Inter-patch dispersal Route redundancy Route vulnerability Habitat availability Component of Connectivity Table 5.1: Summary of the scale-specific classification framework for network statistics that quantify habitat connectivity. Chapter 5. Classification of network-connectivity statistics 105 transformations can be performed among single-valued statistics and distributions for different levels in the network (Brinkmeier and Schank, 2005). Transforming network statistics and distributions from low to higher levels of analysis involves globalization or scaling-up the dependence on elements to neighborhoods to clusters to the entire network. A local-level statistic pertaining to a single network element, such as the de- gree of a node, could undergo globalization to transform it to a network-level statistic, such as the maximum, minimum, summation, or average degrees of all nodes. To trans- form a low-level statistic into a higher-level distribution involves counting the number of elements, neighborhoods, or clusters that lie in a particular range of values (e.g., ab- solute, relative or cumulative degree distribution). Converting distributions to single- valued statistics involves parameter elimination whereby a single value is calculated from a sequence of values given by a distribution. An example of parameter elimina- tion would involve calculating the power law exponent of a degree distribution. Clearly, researchers and practitioners must decide the level at which connectivity statistics are to be calculated for their particular study and their analysis may involve one or more of the aforementioned transformations.

5.6.2 Components of habitat connectivity

For the purpose of this classification framework, I have identified three important com- ponents of habitat connectivity that are measured by the different habitat-network- connectivity statistics: (1) inter-patch dispersal, (2) route redundancy, and (3) route vulnerability. These are not direct measures of functional habitat connectivity per se, but they provide important information on the different components of the species- landscape interaction that result in functional habitat connectivity. The final compo- nent of connectivity that I identified, habitat availability, is more of an outcome than a cause of connectivity, but I have included it nonetheless because it is commonly mea- sured in connectivity assessments of habitat networks. Chapter 5. Classification of network-connectivity statistics 106

Figure 5.2: Examples of different levels of analysis within a habitat network. Habitat patches (grey) are connected by multi-route links (black). Chapter 5. Classification of network-connectivity statistics 107

The first component of connectivity is inter-patch dispersal likelihood which ac- counts for movements among habitat patches. Inter-patch dispersal depends on the source strength of donor habitat patches, the attraction strength of recipient habitat patches, and the probability of dispersal through the matrix (Hanski, 1999b; Tischen- dorf and Fahrig, 2000b). Network-connectivity statistics range from simple measures of habitat-patch source strength, such as “node out-degree” (Schick and Lindley, 2007;

Treml et al., 2008), to integrated measures of source strength and dispersal probability, such as ”dispersal flux” (Urban and Keitt, 2001). Inter-patch dispersal occurs across a multitude of spatial and temporal scales (Mueller and Fagan, 2008) which requires the use of scale-specific measures based on levels of analysis within the network. For example, at the scale of movement between a pair of habitat patches or the element level, “dispersal flux” focuses on the recruitment potential of each patch (estimated by their areas or quality-weighted areas) and the probability of dispersal between the two patches (based on a dispersal kernel and/or properties of the intervening matrix).

At the neighborhood level, “connectivity correlation” measures the degree of compart- mentalization and identifies neighborhoods with high levels of mixing. At the cluster level, “diameter” measures the inter-patch distance an organism would have to traverse to span the cluster. At network level, the “Gamma index” measures the number of di- rectly connected pairs of habitat patches in the network.

The second component of habitat connectivity is route redundancy which measures the presence of multiple movement routes among habitat patches. Including multi- ple movement routes into connectivity analyses acknowledges that individuals rarely use a single optimal route due to variability in behavior and perception (Wiens, 2001;

B´elisle,2005; Driezen et al., 2007). Multiple routes may also be used either because the theoretical optimal route is unattainable or because there exist multiple ways to disperse optimally (or near optimally) among habitat patches in a landscape (Wiens,

2001). An example of the latter is provided by Farmer and Wiens (1998) who mod- Chapter 5. Classification of network-connectivity statistics 108 eled long-distance migration of shorebirds and showed variability in individuals’ migra- tion patterns even when all individuals were making optimal decisions about when and where to refuel. Route redundancy is desirable to maintain in a habitat network from a conservation-planning perspective because it produces more stable connections among habitat patches in the event that natural or anthropogenic disturbances compromise one or more routes (Moilanen et al., 2006; Urban et al., 2009). Redundant movement routes between a pair of habitat patches may be quantified using statistics from cir- cuit theory which specialize in measuring cumulative flux between habitat patches dis- tributed over multiple connections (routes). “Resistance distance” is one such measure

(McRae, 2006; McRae and Beier, 2007; McRae et al., 2008); it measures the opposi- tion to movement of a set of dispersal routes between a pair of habitat patches and de- creases with the addition of more movement routes. Redundancy at neighborhood level can be investigated with the “clustering coefficient” which is a measure of the fraction of triangles present in the network. Triangular network topology means that for each pair of directly connected nodes, there exists at least one more indirect path connect- ing the nodes through a mutual neighbor. A similar measure at the network level is

“meshedness” which compares the actual number of links in the network to the num- ber in the corresponding triangulated planar graph (to which no additional links can be added without creating a non-planar graph).

The third component of habitat connectivity is route vulnerability which describes the degree to which the landscape structure funnels or scatters the movements of a particular species (Lees and Peres, 2008; Pinto and Keitt, 2009). When movements are funneled through particular locations on the landscape then maintaining habitat connectivity hinges on protecting those areas as movement passageways. The pres- ence of such dispersal bottlenecks makes the network more vulnerable to being dis- connected due to habitat destruction, particularly if habitat is destroyed in a man- ner that preferentially destroys those areas (i.e., low attack tolerance; Albert et al., Chapter 5. Classification of network-connectivity statistics 109

2000). An element-level analysis of vulnerability can identify areas (grid cells) in the intervening matrix surrounding habitat patches through which dispersers are likely to move when traveling from one habitat patch to another (i.e., “current density”; Table

5.3). Pinto and Keitt (2009) showed that dispersal bottlenecks in the matrix emerge when the quality of different landcover types in the matrix is spatially autocorrelated and hence there are large clumps of high-quality matrix landcover types to attract dis- persers. At higher levels of analysis, vulnerability can be measured by the presence of

“cut nodes” and “cut links” which act as funnels for movements through the network.

If these nodes or links are removed they will disconnect a cluster to create two smaller clusters and thereby decrease connectivity at both the cluster and network levels.

The final component of habitat connectivity, habitat availability, considers the area inside a patch of habitat as a space where connectivity occurs in addition to inter- patch movements (Hanski, 1999a; Pascual-Hortal and Saura, 2006; Saura and Pascual-

Hortal, 2007). The problem with measures of habitat connectivity based solely on inter- patch movements is that they estimate zero connectivity in landscapes with a sin- gle habitat patch (regardless of its size; Tischendorf and Fahrig, 2000b). The habitat availability component recognizes that habitat connectivity may result either from a spatial configuration of habitat patches that promotes inter-patch movements or from large tracts of contiguous habitat (Ferrari et al., 2007). Pascual-Hortal and Saura

(2006) have proposed that conservation be guided by indices of habitat availability that integrate the amount of habitat and inter-patch connectivity into a single measure.

Network-connectivity statistics often include information about the area of habitat patches at different levels of analysis such as node area and cluster area. At the net- work level, habitat availability can be measured as the maximum or expected area of habitat connected in clusters or through the use of more integrated measures such as

“integral index of connectivity” (Pascual-Hortal and Saura, 2006) or “probability of connectivity” (Saura and Pascual-Hortal, 2007). These network-level measures of habi- Chapter 5. Classification of network-connectivity statistics 110 tat availability will be low either if habitat patches are poorly connected by inter-patch movements or if habitat patches are well connected but the amount of habitat is low

(Saura and Pascual-Hortal, 2007).

5.7 Missing habitat-network connectivity statistics

This classification framework of network statistics (Table 5.1) reveals a striking im- balance in the distribution of measures across different levels of network analysis and among the different components of habitat connectivity. It is evident that most network- connectivity statistics apply at the level of the entire network or individual elements and indeed these are the network levels that are most commonly measured in habitat- connectivity assessments. Analyses at the level of clusters are also fairly well repre- sented by connectivity statistics and have been identified as an important intermediate scale at which to measure connectivity (Urban, 2005; Urban et al., 2009). For example, the size of connected clusters of habitat patches has been correlated with the spatial distribution of Woodland caribou (Rangifer tarandus caribou; O’Brien et al., 2006).

Few connectivity statistics quantify habitat connectivity at the neighborhood level despite the prominence of this level of analysis in other disciplines applying network theory (dyadic and triadic levels in Wasserman and Faust, 1994; Newman, 2003; Proulx et al., 2005). The importance of neighborhoods in a particular habitat network will likely depend on the distances among habitat patches relative to the perceptual range of a species. The connectivity structure of neighborhoods will presumably be more im- portant if the neighborhood level corresponds more closely to the scale of the sensory information window to which an animal can respond (Lima and Zollner, 1996). For- tuna et al. (2006) provide an elegant example of the importance of the neighborhood- level statistic “clustering coefficient” (Table 5.2) in a dynamic network of temporarily

flooded ponds. In their system, a high clustering coefficient provided the opportunity Chapter 5. Classification of network-connectivity statistics 111 for an amphibian, which had moved from a dry to a flooded pond, to move again to another flooded pond if the conditions of the former were not suitable for .

Another means of quantifying neighborhood topology in networks is through the identification of “network motifs” (Milo et al., 2002; Proulx et al., 2005). Network motifs are small repeated patterns or subgraphs that occur significantly more often than expected from random networks. A wide variety of biotic and abiotic networks have been found to contain motifs, some of which are common across networks derived from , neurobiology, ecology, and engineering (Milo et al., 2002, 2004). For example, seven separate food webs contained a four-node motif in which two species shared a common predator and prey; this motif was also found in C. elegans neuronal network and five technological networks (Milo et al., 2002). The presence, number, and distribution of network motifs has been linked to the functioning of biological networks such as gene regulatory networks (Becskei and Serrano, 2000; Shen-Orr et al., 2002) and transcription networks (Mangan and Alon, 2003). To identify motifs in a network, it is possible to search and count all possible configurations of subgraphs with a fixed number of nodes (e.g., all three- and four-node subgraphs; Milo et al., 2002) or to re- strict the search to specific motifs that are hypothesized to be important for function- ality.

Motifs have yet to be identified in habitat networks. I propose that the presence and number of source-sink motifs could indicate inter-patch dispersal at the neighbor- hood level. This motif is rooted in the Pulliam (1988) source/sink model which has also been invoked as a potential conceptual foundation for habitat networks (Urban et al., 2009). A source-sink motif would be defined as a connected pair of source and sink habitat patches in which source and sink habitat patches have positive and neg- ative net reproduction respectively. To detect source-sink motifs, the habitat network must be delineated with directional links (i.e., as a digraph, Box 5.2) differentiating the outward dispersal from sources and the incoming dispersal into sinks. The long-term Chapter 5. Classification of network-connectivity statistics 112 persistence of populations occupying sink patches in the habitat network should also be linked to the presence of source-sink motifs.

The habitat-network-connectivity framework (Table 5.1) also highlights a dearth of network statistics that measure route redundancy at higher network levels. Recently there has been an emphasis placed on quantifying route redundancy at the element level via the use of circuit theory (McRae et al. 2008). These methods have focused on the redundancy of multi-route links that connect pairs of neighboring habitat patches with a series of routes through the matrix (Theobald, 2006; Pinto and Keitt, 2009).

Redundancy also exists at the level of clusters between nodes that are not neighbors.

Multiple paths of varying lengths indirectly connect non-neighboring nodes by travers- ing along a series of connected intermediate links and nodes (Box 5.2). I propose the use of “path redundancy”, which reflects the number of possible paths between pairs of nodes (Kim et al., 2003), as a measure of redundancy at the cluster level. The redun- dant paths may be calculated and counted by modifying any suitable network search algorithm such as Dijkstra’s well known shortest path algorithm (see Pinto and Keitt,

2009, for an example of how to modify Dijkstra’s algorithm to find redundant routes).

An alternative measure of path redundancy is the sum of the ’redundancy degrees’ of intermediate nodes in the shortest path (Kim et al., 2003). A node’s redundancy de- gree is the number of links incident to that node apart from the incoming and outgoing links involved in the shortest path. A higher number of redundancy degrees would in- dicate a route consisting of intermediate nodes with relatively more neighboring nodes and hence the potential for more redundant paths toward the destination (Kim et al.,

2003). Chapter 5. Classification of network-connectivity statistics 113

5.8 Selecting network connectivity statistics

The network connectivity statistics discussed here (Table 5.1) range from reduction- ist measures of the basic features of a network to integrated measures that simulta- neously measure multiple aspects of network structure. Some of these statistics may be correlated either because they represent the same basic aspect of network struc- ture (e.g., class coincidence probability and landscape coincidence probability; Table

5.3) or because different aspects of the network structure are correlated for the partic- ular landscape under investigation. Presumably, a relatively small subset of all possi- ble network-connectivity statistics would suffice to quantify habitat connectivity for a particular study system, as is the case with other landscape pattern metrics (Riiters et al., 1995; Cain et al., 1997; Cushman et al., 2008). It is unlikely however that the exact same set of network-connectivity statistics will be appropriate for all studies that differ with regards to location, data models (vector/raster), scale (grain and extent), and objectives. The choice of statistics should explicitly reflect a hypothesis about the observed habitat network and how the structure of the habitat network affects key eco- logical processes. This classification framework provides researchers and practitioners who are faced with this choice much needed clarification on the ecological meaning of existing network connectivity statistics and the relevant levels of analysis at which they can be applied.

5.9 Spatio-temporal connectivity in dynamic habi-

tat networks

Most habitat network models to date have assumed that the composition and configu- ration of landscapes is static; hence, most network-connectivity statistics quantify habi- tat connectivity at a snapshot in time. In reality, the amount and spatial configuration Chapter 5. Classification of network-connectivity statistics 114 of habitats and matrix landcover types change through time due to natural processes and anthropogenic pressures. However, these landscape dynamics are implicitly incor- porated in network statistics that quantify route redundancy and vulnerability because they measure the presence of alternate movement routes through the network that can maintain overall connectivity in the event that habitat patches or links are destroyed.

A more explicit accounting of network dynamics involves assessing network “robust- ness” or the sensitivity of network connectivity to the removal of nodes or links (New- man, 2003; Minor and Urban, 2008). A small handful of studies have considered the implications of removing nodes and links on habitat-network connectivity measured as:

(1) network diameter, recruitment potential, and area-weighted dispersal flux by Ur- ban and Keitt (2001); (2) clustering coefficient and node-degree correlation by Fortuna et al. (2006); (3) probability of metapopulation extinction in 100 years by Bode et al.

(2008); (4) network size by Matisziw and Murray (2009). In these studies, nodes or links are removed either randomly or in a priority sequence to emulate natural or an- thropogenic landscape disturbances. More robust networks can tolerate a higher frac- tion of nodes or links being removed before habitat connectivity significantly decreases.

Robustness has been shown to be highly related to node degree (Albert et al., 2000) and is likely also measures route redundancy and vulnerability (see Missing habitat- network statistics).

Although these robustness assessments provide valuable insights into the conse- quences of completely destroying habitat patches or links, they do not capture the full range of variability in patch spatial pattern changes. Habitat spatial features may be categorized into primitive events: generation, disappearance, expansion, shrinkage, union, and division (Sadahiro and Umemura, 2001). These primitive events may hap- pen simultaneously across dynamic landscapes which makes the task of quantifying concurrent changes in habitat connectivity non-trivial. Characterizing habitat con- nectivity in these dynamic landscapes requires an explicit analysis of spatio-temporal Chapter 5. Classification of network-connectivity statistics 115 changes in landscape structure.

I propose the construction of spatio-temporal habitat networks to measure habitat connectivity between two maps of a single landscape at two different times. Spatio- temporal links among the set of habitat patches at time one and the set of habitat patches at time two would have a temporal length equal to the time step between the two landscape maps and a spatial length equal to the spatial distance between end- points (patch centroids). Element- and neighborhood-level statistics could be calcu- lated to detect regions of highest connectivity changes based on classified primitive events to produce corresponding “temporal link types”. Cluster- and network-level statistics could be calculated to estimate the degree of change based on the distribu- tion of temporal link types or the mean spatial distance of the shortest link for each patch.

5.10 Conclusions

The number of network-connectivity statistics presents a daunting task in selecting among them for any particular assessment of habitat connectivity. This classification framework provides a much needed description of the qualitative relationships among network-connectivity statistics. Categorizing statistics according to (1) their level of analysis within the network and (2) the component of habitat connectivity that they quantify allows for an ecologically meaningful choice of statistic grounded in hypothesis testing. One of the major advantages of network thinking is that it encourages ecolo- gists to focus on multiple levels of analysis and ask multi-scale questions (Kotliar and

Wiens, 1990) about how habitat patches are embedded within a habitat network and how the network structure emerges from local connections between habitat patches.

The importance of local patterns on overall landscape connectivity has been stressed in static (Bascompte and Sole, 1996; With and King, 1999) and dynamic (Matlack and Chapter 5. Classification of network-connectivity statistics 116

Monde, 2004; Wimberly, 2006) landscapes. Ultimately the goal of network connec- tivity analyses is to determine how the structure of habitat networks influences the persistence of ecological communities. An important step towards this goal will in- volve determining how well measures of potential connectivity can predict actual func- tional connectivity. Separately quantifying the four components of habitat connectivity at different network levels can help to decompose the relationship between potential and function connectivity to move towards identifying the most important aspects of habitat network structure to maintain. Until these statistics have been tested empiri- cally, composite measures in network theory (e.g., Reachability; Table 5.3) should be avoided. Chapter 5. Classification of network-connectivity statistics 117 (Bodin and Norberg, 2007; Bode et al., 2008;et Brooks al., 2008; Minorban, and 2008; Ur- Treml etUrban al., et 2008; al., 2009) (Brooks et al., 2008;et Urban al., 2009) (Bunn et al., 2000;Keitt, Urban 2001; and Estrada-Pena et al., 2006) (Cantwell and Forman, 1993; Forman, 1995; Marcot, 1998; Ricotta et al., 2000;2003; Keitt, Fortuna et al.,Jord´anet 2006; al., 2007; MinorUrban, and 2007) (Keitt, 2003) (Jord´an,2003; Fortuna et al., 2006; Minor and Urban, 2008) (Franc, 2004; Minor and Urban, 2008) (Ferrari et al., 2007;et Jord´an al., 2007; EconomoKeitt, and 2008; Minor andban, Ur- 2008) Indicates habitat patches that serve as stepping stoneslights and routes high- that aredispersal important pathways Indicates habitat patches that occupy a central positionhabitat in network the due toimity their to prox- other habitatIndicates patches the maximum isolationa of habitat patch and influence of aa habitat habitat patch network in Indicates the potential fortat a patch habi- to linkpatches other habitat Indicates the degree tonode’s which neighborhood a is agraph complete Indicates compartmentalization in a network which mayspread reduce of the a cascading disturbance habitat patches in thedistinguishes cluster between and a single, large patchy populations (small value) and a metapopulationvalue) (large Table 5.2: Topological connectivityapplied statistics to that quantify have connectivity been in habitat networks. Number of nodes inbetween the a geodesic focal path nodeminal and node the nearest ter- Average degree of athe focal average node degree relative of to its neighbors Mean geodesic distance in a cluster Indicates the reachability of all Number (or proportion) ofgeodesic pairwise paths in thethrough network a that focal pass node Mean geodesic distance ofevery a other focal reachable node node and in the network Largest geodesic distance betweennode that and any other node Average fraction of adiately focal connected node’s neighbors imme- thatimmediately are connected also neighbors with each other Node degree Number of links incident on a focal nodeNode depth (topological) Indicates the potential accessibility Connectivity correlation path length (topological) centrality Closeness centrality (topological) Eccentricity (topological) coefficient ScaleElement Betweeness Name DefinitionNeighborhood Clustering Ecological Significance References Table 5.2 continued on following page. *Note that geodesic paths and distances are calculated and measured in topological distances. Cluster Characteristic Chapter 5. Classification of network-connectivity statistics 118 (Bunn et al., 2000;and Halpin Bunn, 2000; UrbanKeitt, and 2001; Fagan, 2002; Manseau et al., 2002;Hortal Pascual- and Saura, 2006; Jord´anet al., 2007; Brooks et al., 2008) (Marcot and Chinn, 2003; 1982; Jord´an, Pascual-Hortal and Saura, 2006; SauraPascual-Hortal, and 2007; Schick and Lindley, 2007; Saura, 2008; Treml et al., 2008) (Calabrese and Fagan, 2004; Brooks, 2006; Ferrari et2007) al., (Ricotta et al., 2000;2001, Jord´an, 2003; Pascual-Hortal and Saura, 2006; Saura, 2008) (Marcot and Chinn, 1982; Cantwell and Forman, 1993; Ricotta et al., 2000;2001; Jord´an, Urban and Keitt,Minor 2001; and Urban, 2007; Treml et al., 2008) (Bodin and Norberg, 2007; Treml et al., 2008) (Ricotta et al., 2000;2001, Jord´an, 2003) (Fagan, 2002) patches that are mutuallyable reach- cluster based on thegeodesic maximum distance between anyof pair habitat patches inIndicates the smaller cluster geodesic distances between habitat patches Indicates the presence ofto barriers dispersal if habitatdestroyed patches are Indicates the presence ofthat key hold links clusters ofgether habitat to- Indicates smaller geodesic distances between habitat patches Indicates the variation inber the of num- habitat patches per cluster immediately connected habitat patches in the cluster Sum of the geodesicThe distances Wiener in index a may cluster. the be corresponding normalized chain by andworks planar (i.e. net- the leastnetworks and respectively most that connected areorder) of the same Maximum geodesic distance in a cluster IndicatesSum the of compactness of the inverse a tances of in the a geodesic cluster. dis- be The normalized Haray in index aWiener may similar index fashion to the Number of nodes whosenects removal the discon- cluster by creating a new cluster Number of links whosethe removal cluster disconnects by creating a new cluster Standard deviation of clusterby order mean divided cluster order Wiener index and Normalized Wiener index Cluster size Total number of links in the clusterDiameter (topological) Indicates theHarary number index of pairs of and Normalized Harary index Number of cut-nodes Number of cut-links variation (CV) of cluster order ScaleCluster Name Cluster order Number of nodes in the cluster Definition Indicates the number of habitat Ecological Significance References Table 5.2 continued on following page. *Note that geodesic paths and distances are calculated and measured in topological distances. Network Coefficient of Chapter 5. Classification of network-connectivity statistics 119 Jr´n 2003) (Jord´an, (Franc, 2004; Brooks, 2006; Fortuna et al., 2006;et Rhodes al., 2006; Minor2007; and Brooks Urban, et al.,Economo 2008; and Keitt, 2008; Minor and Urban, 2008) (van Langevelde et al.,Franc, 1998; 2004; Pascual-Hortal and Saura, 2006; Bodin and Norberg, 2007; Saura and Pascual-Hortal, 2007; Economo and Keitt, 2008;nor Mi- and Urban, 2008;et Treml al., 2008) (Forman, 1995; Ricotta et2000; al., Jord´an,2001, 2003) (Marcot and Chinn, 1982; Fagan, 2002; Jord´an,2003; Treml et al., 2008) (Marcot and Chinn, 2003; 1982; Jord´an, Pascual-Hortal and Saura, 2006; SauraPascual-Hortal, and 2007; Schick and Lindley, 2007; Treml et al., 2008; Saura2009; and Matisziw Torne, and Murray, 2009) (Bunn et al., 2000;and Halpin Bunn, 2000; UrbanKeitt, and 2001; Jord´an,2003; Pascual-Hortal and Saura, 2006; Saura and Pascual- Hortal, 2007; Saura, 2008) of immediately connected habi- tat patches standardized bynumber the of habitat patchesnetwork in the Indicates the distribution oftential po- source and sinkpatches habitat as well asof the the robustness network todestruction habitat patch Indicates the compactness ofhabitat a network based onimum the distance max- between anyhabitat pair patches of Indicates the number ofately immedi- connected habitat patches clusters of habitat patches patches in the habitat network immediately connected habitat patches Probability distribution of nodeover degrees the entire network Maximum geodesic distance incluster the largest the number of linksplanar in network the corresponding Total number of connected sub-graphs Number of distinct, unconnected Degree distribution Diameter of largest cluster (topological) Gamma index Number of links inNetwork network order divided by Total number ofNetwork nodes size within network Total number of links Indicates within the network number of habitat Number of Indicates theclusters number of pairs of ScaleNetwork Connectance Name Density of links in network Definition Indicates the number of pairs Ecological Significance References Table 5.2 continued on following page. *Note that geodesic paths and distances are calculated and measured in topological distances. Chapter 5. Classification of network-connectivity statistics 120 (Brooks et al., 2008) (Marcot and Chinn, 1982; Cantwell and Forman, 1993; Forman, 1995; Jord´an,2003; Ferrari et al., 2006) (Forman and Gordon, 1986; Forman, 1995) (Bunn et al., 2000;and Halpin Bunn, 2000; UrbanKeitt, and 2001; Fagan, 2002; Manseau et al., 2002;2003; Jord´an, Pascual-Hortal and Saura, 2006; Brooks et2008) al., (Marcot and Chinn, 1982; Cantwell and Forman, 1993; Ricotta et al., 2000;2001; Jord´an, Urban and Keitt,Minor 2001; and Urban, 2007; Treml et al., 2008) (Bodin and Norberg, 2007; Treml et al., 2008) pairs of immediately connected habitat patches in theIndicates cluster the average orof variation the accessibility ofpatch each habitat Indicates the presence ofnate alter- pathways which may assist organisms to avoid disturbancepredation and Indicates the maximum ormum mini- number of habitatthat patches are mutually reachable Indicates the presence ofto barriers dispersal if habitatdestroyed patches are Indicates the presence ofthat key hold links clusters ofgether habitat to- Number of connected nodesor in smallest the cluster largest Number of nodes whosenects removal the discon- network byter creating a new clus- Number of links whosethe removal network disconnects by creating a new cluster Mean number of links per clusterMean or standard deviationof of links the per number node IndicatesActual the number average of number loops of by in number network of divided loopsmaximal/triangulated in planar the network corresponding Order of largest or smallest cluster Total number of cut-nodes Total number of cut-links Size Mean or stan- dard deviation of node degree Meshedness or Alpha Index or Network Circuity ScaleNetwork Mean Name Cluster Definition Ecological Significance*Note that geodesic paths and distances are calculated and measured in topological References distances. Chapter 5. Classification of network-connectivity statistics 121 (McRae et al., 2008) (Urban et al., 2009) (McRae et al., 2008) (Urban and Keitt, 2001; Minor and Urban, 2008) (Bunn et al., 2000;and Urban Keitt, 2001; Urban2009) et al., Indicates the expected timeindividual for to an move fromtat one patch habi- to theagain other if and they back move randomly Indicates the areas inscape the which land- dispersers areto likely pass through whenone moving habitat from patch to another Indicates the amount ofbetween movement a pair of habitat patches Indicates the maximum isolationa of habitat patch Indicates habitat patches that occupy a central positionhabitat in network the due toimity their to prox- other habitat patches , and y , xy and 1 R x n is the resistance =1 y X xy n =1 R X x , v uv ˆ R = and : v u uv is the effective resistance be- CT and uv Table 5.3: Node- andhave link-weighted been connectivity applied statistics to that quantify connectivity in habitat networks. ˆ R u quality-weighted area of nodesthe adjacent link to are multipliedof by dispersal the between probability nodes Largest geodesic distance betweennode that and any other node of the link connectingn nodes is the number of nodes in the network resentation of a habitatmeasures network the that flow ofthrough current a (dispersers) network whencurrent one (a ampere fixed of numberinjected of into dispersers) one is nodeconnected and to the ground other (are is through forced the to network). move tween nodes A property of thenodes set of links link between where Mean geodesic distance ofevery a other focal reachable node node and in the network ) CT Dispersal flux A property of a link in whichEccentricity the area or (metric) Current density A property of each cell in a raster rep- centrality (metric) Commute time ( ScaleElement Closeness Name Definition Ecological Significance References Table 5.3 continued on following page. *Note that geodesic paths and distances are calculated and measured in metric distances. Chapter 5. Classification of network-connectivity statistics 122 (Magle et al., 2009) (Fortuna et al., 2006) (Schick and Lindley, 2007; Brooks et al., 2008;et Treml al., 2008) (Miller and Russell, 2004;nor Mi- and Urban, 2007;and Schick Lindley, 2007) (McRae, 2006; McRae and Beier, 2007; McRae et2008) al., Indicates the role ofpatch a in habitat connecting otherpatches habitat through itself Indicates the amount ofment move- into a habitatprimary patch neighbors from its Indicates the connectivity ofhabitat a patch to allpatches other in habitat the habitatassuming network, that patch sizespermeability and of the the intervening matrix are important ity Multiple links between anodes pair represent of alternative move- ment routes and effectivedecreases resistance with the additionlinks of more k A v and , u ) i , ¯ A and all other A ( uvi uvi k k R R A ik m m =1 =1 Y i i X w is the number of links in = n =1 X k m uv = : . ˆ R v v i k is the resistance (or opposition) T offers to movement between is the area of the focal node, i and and uvi i is the mean area of all nodes, and and R A u i u ¯ A is a term that describes the potential ik that link where nodes the set of links between nodes where A property of aof node the which area- is and aevery cost-weighted function other distance node to in the network: is the area ofarea, a single node in the study A property of thenodes set of links between Pearson product-moment correlation coef- ficient between the directedundirected degree degree and of the a node cent to a node patches based on ancosts estimate derived of from movement multiplenodes links between a w movement between patch ) ˆ R ) i T resistance/ Resistance distance ( Node degree correlation Node in-degree Total number of linksNode coming influx into a node Indicates Sum habitat of patch fluxes accessibil- for all incoming links adja- Modified inci- dence function model ( ScaleElement Effective Name Definition Ecological Significance References Table 5.3 continued on following page. *Note that geodesic paths and distances are calculated and measured in metric distances. Chapter 5. Classification of network-connectivity statistics 123 Jr´n 2003,(Jord´an, modified slightly in Jord´anet al.,Vasas 2007; et al., 2009) (Bode et al., 2008) (Jord´anet al., 2007; Minor and Urban, 2008) 2007; Brooks et al.,Treml 2008; et al., 2008) (Miller and Russell, 2004;nor Mi- and Urban, 2007;and Schick Lindley, 2007) (Minor and Urban, 2007) (Ferrari et al., 2007) (van Langevelde et al.,Bunn 1998; et al., 2000;Keitt, Urban 2001; and Brooks, 2006; Ferrari et al., 2007;and Minor Urban, 2008; Treml2008) et al., Indicates topological, link- weighted, and node-weighted prop- erties that affect theof connectivity a node Indicates the rapidity withthe which average unoccupied patchnetwork in can a be recolonized habitat patches in thedistinguishes cluster between and a single, large patchy populations (small value) and a metapopulationvalue) (large that is accessible toin an that individual cluster Indicates the amount ofout movement of a habitatprimary patch neighbors from its Indicates the reproductive poten- tial of a habitat patch Indicates the compactness ofcluster a based on themetric maximum distance geo- between anyhabitat pair patches of in the cluster ) i ( is i , )) i CC ( ) is its i , ( i LP Smax ) i davtgr CC LP Smax − i ) + i D ( ( davtgr ( is the degree of node i = i D I average geodesic distance, and is the maximal connectedsize local after population its deletion Mean path strength overpatches all in pairs a of cluster Mean geometric distance in network Indicates the reachability of all where its clustering coefficient, Total number of links leaving a nodecent to a node Indicates habitat patch influenceNode area (i.e. habitattiplied patch by size) node mul- quality (Schickaverage (defined and distance as Lindley, between the eacha pixel habitat in patch andvalue a ranges non-habitat from edge; 0-1) Maximum geodesic distance betweenin nodes a cluster ) i I strength Characteristic path length (metric) Cluster Area Sum of areas of all nodes in cluster Indicates the amount of habitat degree Node outflux Sum ofQuality- fluxes for all outgoingweighted links area adja- Reachability/ Importance index ( Diameter (metric) Cluster Average path ScaleElement Node Name out- Definition Ecological Significance References Table 5.3 continued on following page. *Note that geodesic paths and distances are calculated and measured in metric distances. Chapter 5. Classification of network-connectivity statistics 124 (Bode et al., 2008) 2000) (Jord´an, (Ferrari et al., 2007) (Bode et al., 2008) (Saura, 2008; Pascual-Hortal and Saura, 2008; SauraPascual-Hortal, and 2007; Pascual- Hortal and Saura, 2006) Indicates the probability thatrandomly two located points (orviduals) indi- within the habitatto belong the same cluster Indicated the probability ofment move- between a pairpatches of habitat Indicates the proportion ofavailable habitat to an individualhaving without to traverse theIncreasing matrix asymmetry has aative neg- effect on themetapopulation probability persistence of Indicates the probability that habitat patch immigration and emigration will operate successfully in the habitat network , i i c , 2 ) i C c A ( =1 is the probability that i NC X P = CCP is the number of clusters, , where P is the sum of all node areas in the − NC C A = 1 network and with a path betweenhas two the patches maximum which associatedmovement probability of R the area of the largest cluster connectivity pattern and itsequivalent symmetric (the mean ofconnections the in asymmetric both directions) where the habitat network becomesinto disconnected two or more isolated clusters is the sum of the node areas in cluster ) The difference between the asymmetric ) Z R ) CCP Reliability ( Asymmetry ( Class Coinci- dence Probabil- ity ( ScaleCluster Name Path strength The probability of movement associated DefinitionNetwork Area index Area of largest habitat patch divided by Ecological SignificanceTable 5.3 continued on following page. *Note that geodesic paths References and distances are calculated and measured in metric distances. Chapter 5. Classification of network-connectivity statistics 125 (Keitt et al., 1997;Pena, Estrada- 2002; Calabrese and Fagan, 2004; Rothley, 2005) (Miller and Russell, 2004) (Manseau et al., 2002;et James al., 2005; O’Brien2006; et Fall al., et al., 2007) Indicates the average distanceindividual an is capable ofbefore dispersing reaching a barrier,randomly if on placed the landscape Indicates the probability ofcessful suc- dispersal among habitat patches based on themultiple presence dispersal of routes, variation in matrix quality, andbarriers dispersal Indicates the mean areathat of a habitat disperser hasthe access mean to cluster (i.e. size),a if random placed habitat in patchnetwork within the , 2 ) s are n . i s y < y > − and i i y x , are the centroid s + ( s 2 R ) s n n =1 m s X =1 m s X < y > < x > = − i and C x ( p is the number of clusters and n =1 i X < x > m 1 n = is the radius of gyration calculated as R R follows: where is the number of grid cells in cluster The area-weighted mean clusterthat size. size Note here refersnodes to rather the than number number of cluster of links in the the coordinates of theof surrounding the polygon patch. Thispatch corrects shapes for irregular where coordinates of the patch and Mean connectivity of allconnectivity nodes of where each the nodeof is its a area, function itsmovement effective route distances from of all each its other barriers nodes, along and all dispersal routes Expected Cluster Size Ecologically scaled connectivity length ScaleNetwork Correlation Name Definition Ecological Significance References Table 5.3 continued on following page. *Note that geodesic paths and distances are calculated and measured in metric distances. Chapter 5. Classification of network-connectivity statistics 126 (Bunn et al., 2000;Keitt, Urban 2001; and Estrada-Pena et al., 2006; Minorban, and 2007; Ur- Pascual-Hortal and Saura, 2007; SauraPascual-Hortal, and 2007; Saura, 2008) (Pascual-Hortal and Saura, 2006; Saura and Pascual- Hortal, 2007; Saura, 2008) (Jord´an,2003; Bodin and Norberg, 2007; Jord´anet al., 2007) (Pascual-Hortal and Saura, 2006, 2007; Saura and Pascual-Hortal, 2007; Saura, 2008) ) among them ij nl Indicates the maximum metapop- ulation size within thenetwork habitat Indicates the probability thatrandomly two located points (orviduals) indi- within the landscapepoints (i.e., can lie eithernon in habitat habitat areas) or belongsame to habitat the cluster Indicates total movement among habitat patches in thenetwork habitat Indicates the probability thatanimals two randomly placed within the landscape fall intoeas habitat that ar- are reachableother from (interconnected) each given aof set n habitat patchesconnections and ( the binary is i , and c i , ij , j 2 nl a ) i L i a c 2 L A 1 + ( A n =1 =1 j X i NC X n = =1 i X are the areas of the nodes is the number of links j = ij a LCP nl is the number of clusters, IIC and and i NC a j is the total landscape area is the total landscape area L L and Sum of population sizestional or areas areas in or the propor- largest network cluster where the sum of theA node areas in cluster Sum of flux associatedthe with network each multiplied link by in of the total proportion habitat area in its incident node(s) i where in the geodesic pathgeodesic (i.e. distance) the between topological nodesA i and j. ) ) IIC LCP Maximal con- nected local population size Expected dis- persal flux or Area-weighted flux of connectivity ( Landscape Coincidence Probability ( Scale Name Definition Ecological Significance References Table 5.3 continued on following page. *Note that geodesic paths and distances are calculated and measured in metric distances. Network Integral index Chapter 5. Classification of network-connectivity statistics 127 (Pascual-Hortal and Saura, 2007; Saura and Pascual- Hortal, 2007; Saura, 2008) (Urban and Keitt, 2001; Estrada-Pena et al., 2006) (Bunn et al., 2000;Keitt, Urban 2001; and Estrada-Pena et al., 2006) ) among ij p Indicates the potential toindividuals add to new the populations living in habitat patchesnetwork in the Indicates the potential fordistance long rescue within thenetwork habitat them Indicates the probability thatanimals two randomly placed within the landscape fall intoeas habitat that ar- are reachableother from (interconnected) each given a set of n habitatweighted patches connections and ( the i , ij ? p j a i 2 L a A n =1 j X n =1 is the total landscape i are the areas of the nodes X is the maximum product j L = a ij A ) belonging to each step in ?p ij PC p and i , and a j j and where in the network graph i area. The product probabilityis of the a product path ofabilities all ( the dispersal prob- probability of all paths between patches that path. and ) PC Traversability Diameter of the largest component of the Probability of connectivity ( Scale Name DefinitionNetwork Recruitment Sum of quality-weighted area for all nodes Ecological Significance*Note that geodesic paths and distances are calculated and measured in metric distances. References Chapter 6

Conclusions and future research directions

6.1 Thesis summary

Natural habitats are rapidly being destroyed and fragmented at a global scale (Foley et al., 2005). Conservation measures are required for plant and animal species to sur- vive in the face of this widespread habitat destruction and fragmentation. Contempo- rary conservation biology favours an approach of prioritizing habitats for protection in ecological reserves and conserving or restoring natural levels of connectivity (species’ movements) among them (Margules and Pressey, 2000; Crooks and Sanjayan, 2006).

New methods and strategies presented in this dissertation provide novel ways to in- tegrate important ecological considerations, such as landscape dynamics, species in- teractions, and spatial heterogeneity of habitat types, into reserve selection and habi- tat connectivity assessments. My research has shown that explicitly accounting for these ecological complexities in conservation planning and assessment can result in a re-prioritization of habitats and spatio-temporal land uses to preserve biodiversity.

The need to incorporate stochastic landscape processes, such as succession and

128 Chapter 6. Conclusions 129 natural disturbance regimes, into terrestrial reserve-selection strategies has only re- cently been discussed in either theoretical (Cumming et al., 1996; Bengtsson et al.,

2003; but see Pickett and Thompson, 1978) for an early discussion on this topic) or practical terms (Rayfield et al., 2008; Leroux et al., 2007a,b). One of the first strate- gies proposed to select reserves in dynamic landscapes was to create a system of dy- namic reserves which are re-located through time to maintain representation and per- sistence targets (Cumming et al., 1996). In Chapter 2, I coupled a reserve-selection procedure with a spatially-explicit boreal forest dynamics model to compare the abil- ity of dynamic and static reserves to maintain spatial habitat requirements of Amer- ican Marten through time. This is an underused approach for evaluating alternative reserve selection strategies using a spatially-explicit dynamic simulation model (but see Leroux et al., 2007a,b) despite its prevalence in natural resource management (Fall et al., 2004). Simulation results showed that dynamic reserves moderately improved upon static reserves by maintaining more marten home ranges that were of compa- rable quality to those in static reserves. The performance of the dynamic reserves is thought to be highly contingent on implementation details such as the length of the re-planning interval and the logging regime applied in the surrounding matrix. This study provides a valuable illustration of why comprehensive conservation plans should not rely exclusively on the use of reserves but should also focus on the spatial and tem- poral arrangement of land uses in the surrounding matrix to provide opportunities for re-planning in the future (Lindenmayer and Franklin, 2002). The performance of dy- namic reserves was constrained largely due to inappropriate matrix planning. A focus on complementary reserve and matrix planning will require a novel management prac- tice that adopts a longer time frame. As time progresses, spatial legacies of previous management actions can “lock in” spatial patterns that limit options for managers in the future (James et al., 2007).

Conservation planning based on the requirements of an indicator species, such as Chapter 6. Conclusions 130 the American Marten in Chapter 2, implicitly assumes that by maintaining their habi- tat requirements the conditions will also be suitable for other species (Landres et al.,

1988). There has been considerable debate surrounding the effectiveness of indicator species in conservation planning (Roberge and Angelstam, 2004) with many studies arguing in favour of planning based on the needs of multiple species (Landres et al.,

1988; McLaren et al., 1998; Carignan and Villard, 2002; Thompson, 2006). In Chap- ter 3, I present a multi-species, reserve-selection method that explicitly incorporates consumer-resource interactions. This novel method protects consumer-resource inter- actions by prioritizing areas that are important to maintain connectivity between their distributions. The inclusion of interacting-species connectivity requirements resulted in the spatial aggregation of reserves which is recognized as a desirable feature of re- serve from the perspective of minimizing negative edge effects (Fagan et al., 1999) and reducing implementation costs from ecological and economic perspectives (Ball and

Possingham, 2000). Many reserve-selection methods impose an arbitrary penalty (i.e., boundary-length penalty; Ball and Possingham, 2000) to force reserves to be more spa- tially aggregated; however, this was achieved in the reserves in Chapter 3 by modelling explicitly the spatial connectivity requirements of a consumer and its resources. This represents an important shift away from reserve-selection methods that impose an arbi- trary aggregated geometry to reserve-selection methods that induce spatial aggregation via models of species use of space.

Connectivity among consumer and resource distributions was modelled using a 2- dimensional kernel smoothing with a species-specific foraging parameter for the width of the smoothing kernel. This modelling approach, commonly employed in metapopu- lation studies (Moilanen and Nieminen, 2002; Moilanen et al., 2005), assumes that con- nectivity is a function of geographic distances among habitat patches embedded within a homogeneous matrix. However, the landscape matrix is seldom homogeneous with respect to animal movement and species generally do not follow straight-line routes Chapter 6. Conclusions 131 to reach their destination (Adriaensen et al., 2003; Chardon et al., 2003; Verbeylen et al., 2003; Lindenmayer and Franklin, 2002). The ability of species to move among habitat patches (habitat connectivity) is therefore not simply a function of geographic distance but rather it is a function of the landscape heterogeneity and species-specific movement responses to that heterogeneity (Taylor et al., 1993; Tischendorf and Fahrig,

2000b). Quantifying species perceptions of landscape heterogeneity can be achieved by creating a cost surface that represents the ecological costs of movement through each landcover type. This cost surface can serve as the basis for finding least-cost routes between habitat patches and analyzing habitat connectivity of the landscape using a graph or network approach (Bunn et al., 2000; Chetkiewicz et al., 2006; O’Brien et al.,

2006; Theobald, 2006; Fall et al., 2007). There is no consensus on the best approach to estimate movement costs and there is considerable uncertainty associated with these cost surfaces (Cushman et al., 2006). Different studies have performed restricted sensi- tivity analyses in which they examine a limited set of alternative cost values. In Chap- ter 4, I performed a comprehensive sensitivity analysis of least-cost habitat networks to the relative cost values covering the range of cost values reported in the literature.

Least-cost habitat networks proved to be sensitive to relative cost values of the matrix landcover types but the degree of sensitivity differed among landscapes based on their spatial structure. Spatial deviation among routes was highest when the habitat struc- ture was fragmented and when the amount of hospitable landcover in the matrix was in the range of 20-50

Recognizing this sensitivity of least-cost routes in habitat networks to cost values, novel methods have been developed to identify potential movement among habitat patches through low-cost zones comprised of multiple movement routes rather than single least-cost routes (Pinto and Keitt, 2009; Sutherland et al., 2007, Fall et al., in prep). Specifically, methods from graph, network, and circuit theory are currently be- ing applied to quantify habitat connectivity in habitat networks constructed with mul- Chapter 6. Conclusions 132 tiple movement routes between patches (McRae, 2006; McRae and Beier, 2007; McRae et al., 2008; Urban et al., 2009). Over the last decade, these methods have proposed over sixty connectivity statistics that can be used to quantify different aspects of con- nectivity in habitat networks. To clarify the ecological interpretation of these network connectivity statistics, I developed a classification framework. The framework catego- rizes connectivity statistics based on four components of connectivity (i.e., dispersal likelihood, route redundancy, route vulnerability, and habitat area connected) and four network levels (i.e., element, neighbourhood, cluster, and network). The framework revealed a lack of statistics in the categories of “dispersal likelihood among neighbour- ing habitat patches” and “route redundancy at the level of network clusters”. Hence,

I describe network-motif and path-redundancy statistics that fit these categories. This framework is a timely and important contribution because an overall assessment of net- work connectivity is needed rather than focusing exclusively on structural properties at local levels within the network. By following this framework, one can better under- stand network-connectivity statistics so that explicit hypotheses can be tested about the relationship between network connectivity structure and the persistence of biodi- versity in habitat networks.

Throughout this dissertation, I have emphasized ways in which our conceptual un- derstanding of ecological systems can be incorporated into conservation management and assessment. The methods and results that I have presented can help us act strate- gically to protect biodiversity at a time when the human footprint is rapidly expand- ing. My finding that dynamic reserves will benefit from a whole-landscape approach to conservation planning, emphasizes the need for a better understanding of the interac- tion between reserves and matrix management. I stress that within reserves one should account for the spatial habitat requirements of interacting species, such as consumers and resources, as they are the foundation upon which ecological communities are sus- tained. I also emphasize that within the matrix, heterogeneity in habitat and landcover Chapter 6. Conclusions 133 types has important consequences for habitat network connectivity assessments based on least-cost routes. The sensitivity that I found in these assessments calls for a more robust means of incorporating uncertainty about matrix quality into the delineation of multiple movement routes between habitat patches. I therefore explored the potential for graph, network and circuit analyses to provide a more complete characterization of the connectivity structures of habitat networks at many different network levels. Such a network-connectivity classification framework will improve the ability of conservation planners to select connectivity statistics that can identify habitat patches and zones within the landscape matrix for connectivity conservation or restoration.

6.2 Future research directions

Many important questions remain about how reserve design and the connectivity among reserves affects broad-scale ecological processes that support the persistence of biodi- versity. These processes include dispersal (e.g., demographic rescue, infections disease spread, range expansion, exotic species invasions), species interactions (e.g., metacom- munity dynamics, predation, competition, mutualisms), and gene flow (e.g., genetic drift, genetic bottlenecking, inbreeding depression)(Kareiva and Wennergren, 1995;

Moilanen et al., 2005; Crooks and Sanjayan, 2006). Answering these questions em- pirically is challenging, if possible, given the difficulties associated with manipulating whole ecosystems and replicating landscape-scale experiments (Schindler, 1998). In light of these challenges, I propose three complementary approaches that can be used to provide some insights into questions about the consequences of reserve structure for species persistence.

The first approach would involve combining the spatially-explicit landscape sim- ulation modelling method presented in Chapter 2 with spatially-explicit population dynamics models (Turner et al., 1995; Wintle et al., 2005). This modelling approach Chapter 6. Conclusions 134 allows for comparisons among alternative reserve structures and strategies by examin- ing the range of possible persistence outcomes that are simulated under a given set of assumptions and initial conditions. Another approach would involve the use of natural microcosms or experimental model systems (Srivastava et al., 2004) to experimentally test the causal relationship between reserve network structure and the persistence of biodiversity. These experiments could involve manipulating the topological structure of micro habitat networks in a laboratory setting (e.g., Gonzalez and Chaneton, 2002) and relating those structural differences to community patterns of resident microfauna

(e.g. diversity, relative abundance, distribution of patch abundance and occupancy).

A final approach would be to use “mensurative experiments” (McGarigal and Cush- man, 2002) in which existing habitat networks are carefully selected based on contrast- ing structures. For example, a researcher might look for habitat networks that have markedly different levels of compartmentalization and node isolation, such as random and small-world topologies (Bode et al., 2008; Minor and Urban, 2008). While manipu- lative experimentation at a landscape scale is generally not possible, it should be noted that restoration ecology offers an excellent opportunity for landscape-scale, replicated mensurative (or even possibly manipulative) experiments (Palmer et al., 2006).

The vast majority of habitat connectivity assessments to date have measured con- nectivity at a single snapshot in time. Relating this snapshot of connectivity to ob- servations of species’ distributions and occupancy patterns assumes that landscape structure is static and ignores temporal variation caused by disturbance and succes- sional processes. Hence, static habitat network models may have limited predictive power (Wiens, 1997; Matlack and Monde, 2004; Wimberly, 2006). For example, Mat- lack and Monde (2004) and Wimberly (2006) showed that individuals use ephemeral habitat patches as stepping stones to facilitate dispersal, a phenomenon that would not be accounted for by static models. In Chapter 5, spatio-temporal habitat networks are introduced as a means of quantifying habitat connectivity between habitat networks Chapter 6. Conclusions 135 at successive time periods. These spatio-temporal habitat networks could improve our ability to better manage dynamic landscapes to improve long-term connectivity. This novel application of network theory provides fertile grounds for research into dynamic network structures and the consequences for biodiversity conservation. Bibliography

Abrams, P. A., 1987. Indirect interactions between species that share a predator: va-

rieties of indirect effects. In W. Kerfoot and A. Sih, editors, Predation, pages 38–54.

University Press of New England, Hanover.

Abrams, P. A., 1999. The adaptive dynamics of consumer choice. The American Natu-

ralist 153:83–97.

Abrams, P. A., 2007. Habitat choice in predator-prey systems; spatial instability due to

interacting adaptive movements. The American Naturalist 169:581–594.

Acosta, A., I. C. Blasi, M. L. Carranza, C. Ricotta, and A. Stanisci, 2003. Quantify-

ing ecological mosaic connectivity and hemeroby with a new topoecological index.

Phytocoenologia 33:623–631.

Adriaensen, F., J. P. Chardron, G. DeBlust, E. Swinnen, S. Villalba, H. Gulinck, and

E. Matthysen, 2003. The application of ‘least-cost’ modelling as a functional land-

scape model. Landscape and Urban Planning 64:233–247.

Aebischer, N. J., P. A. Robertson, and R. E. Kenward, 1993. Composition analysis of

habitat use from animal radio-tracking data. Ecology 74:1313–1325.

Albert, R. and A.-L. Barab´asi,1999. Diameter of the World-Wide Web. Nature

401:130.

136 BIBLIOGRAPHY 137

Albert, R., H. Jeong, and A.-L. Barab`si, 2000. Error and attack tolerance of complex

networks. Nature 406:378–382.

Andr´en,H., 1994. Effects of habitat fragmentation on birds and mammals in land-

scapes with different proportions of suitable habitat: a review. Oikos 71:355366.

Andr´en,H., 1999. Habitat fragmentation, the random sample hypothesis and critical

thresholds. Oikos 84:306–308.

Angelstam, P., S. Boutin, F. Schmiegelow, M.-A. Villard, P. Drapeau, G. Host,

J. Innes, G. Isachenko, T. Kuuluvainen, M. Monkkonen, J. Niemela, G. Niemi, J.-

M. Roberge, J. Spence, and D. Stone, 2004. Targets for boreal forest biodiversity

conservation - a rationale for macroecological research and adaptive management.

Ecological Bulletin 51:487–509.

Araujo, M. B., M. Cabeza, W. Thuiller, L. Hannah, and P. H. Williams, 2004. Would

climate change drive species out of reserves? An assessment of existing reserve-

selection methods. Global Change Biology 10:1618–1626.

Arponen, A., R. K. Heikkinen, C. D. Thomas, and A. Moilanen, 2005. The value of

biodiversity in reserve selection: representation, species weighting, and benefit func-

tions. Conservation Biology 19:2009–2014.

Arponen, A., H. Kondelin, and A. Moilanen., 2007. Area-based refinement for selection

of reserve sites with the benefit-function approach. Conservation Biology 21:527–

533.

Baker, W. L. and Y. Cai, 1992. The r.le programs for multiscale analysis of landscape

structure using the GRASS geographical information system. Landscape Ecology

7:291–302. BIBLIOGRAPHY 138

Ball, I. R. and H. P. Possingham, 2000. MARXAN (V1.8.2): Marine Reserve Design

Using Spatially Explicit Annealing. Technical report, A manual prepared for the

Great Barrier Reef Marine Park Authority.

Barab´asi,A. L., 2002. Linked: the new science of networks. Perseus Publishing, Cam-

bridge MA, USA.

Barton, N. H., 1992. The genetic consequences of dispersal. In N. C. Stenseth and

W. Z. J. Lidicker, editors, Animal Dispersal: Small Mammals as a Model, pages 37–

59. Chapman & Hall, London.

Bascompte, J., 2007. Networks in ecology. Basic and Applied Ecology 8:485–490.

Bascompte, J. and R. V. Sole, 1996. Habitat fragmentation and extinction thresholds

in spatially explicit models. Journal of Animal Ecology 65:465–473.

Baskent, E. Z. and S. Keles, 2005. Spatial forest planning: a review. Ecological Model-

ing 188:145–173.

Baskett, M. L., F. Micheli, and S. A. Levin, 2007. Designing marine reserves for inter-

acting species: insights from theory. Biological Conservation 137:163–179.

Beazley, K., L. Smandych, T. Snaith, F. Mackinnon, J. R. Peter Austen-Smith, and

P. Duinker, 2005. Biodiversity considerations in conservation system planning: map-

based approach for Nova Scotia, Canada. Ecological Applications 15:2192–2208.

Becskei, A. and L. Serrano, 2000. Engineering stability in gene networks by autoregula-

tion. Nature 405:590–593.

Beisner, B., D. Haydon, and K. Cuddington, 2003. Alternative stable states in ecology.

Frontiers in Ecology and the Environment 1:376–382.

B´elisle,M., 2000. Influence de la composition et de la configuration du paysage sur le

movement des oiseaux forestiers. Ph.D. thesis, Universit Laval. BIBLIOGRAPHY 139

B´elisle,M., 2005. Measuring landscape connectivity: the challenge of behavioral land-

scape ecology. Ecology 86:1988–1995.

Bengtsson, J., P. Angelstam, T. Elmqvist, U. Emanuelsson, C. Folke, M. Ihse,

F. Moberg, and M. Nystrom, 2003. Reserves, resilience and dynamic landscapes.

Ambio 32:389–396.

Bergeron, Y., S. Gauthier, V. Kafka, P. Lefort, and D. Lesieur, 2001. Natural fire

frequency for the Eastern Canadian Boreal Forest: consequences for sustainable

forestry. Canadian Journal of Forest Research 31:384–391.

Bergeron, Y., B. Harvey, A. Leduc, and S. Gauthier, 1999. Forest management guide-

lines based on natural disturbance dynamics: stand- and forest-level considerations.

Forestry Chronicle 75:49–54.

Bergeron, Y., A. Leduc, B. Harvey, and S. Gauthier, 2002. Natural fire regime: a guide

for sustainable management of the Canadian Boreal Forest. Silva Fennica 36:81–95.

Bergeron, Y. D., D. Cyr, C. R. Drever, M. Flannigan, S. Gauthier, D. Kneeshaw,

E. Lauzon, A. Leduc, H. Le Goff, D. Lesieur, and K. Logan, 2006. Past, current and

future fire frequency in Qu´ebec’s commercial forests: implications for the cumulative

effects of harvesting and fire on age-class structure and natural disturbance based

management. Canadian Journal of Forest Research 36:2737–2744.

Bode, M., K. Burrage, and H. P. Possingham, 2008. Using complex network metrics to

predict the persistence of metapopulations with asymmetric connectivity patterns.

Ecological Modelling 214:201–209.

Bodin, O. and J. Norberg, 2007. A network approach for analyzing spatially structured

populations in fragmented landscapes. Landscape Ecology 22:31–44. BIBLIOGRAPHY 140

Borgatti, S. P., A. Mehra, D. J. Brass, and G. Labianca, 2009. Network analysis in the

social sciences. Science 323:892–895.

Bossenbroek, J. M., H. H. Wagner, and J. A. Wiens, 2005. Taxon-dependent scaling:

beetles, birds, and vegetation at four North American grassland sites. Landscape

Ecology 20:675–688.

Bowman, J., G. Forbes, and T. Dilworth, 2001. Landscape context and small-mammal

abundance in a managed forest. Forest Ecology and Management 140:249–255.

Bowman, J., J. A. G. Jaeger, and L. Fahrig, 2002. Dispersal distance of mammals is

proportional to home range size. Ecology 83:2049–2055.

Bowne, D. R. and M. A. Bowers, 2004. Interpatch movements in spatailly structured

populations: a literature review. Landscape Ecology 19:1–20.

Boyce, M. S. and L. L. McDonald, 1999. Relating populations to habitats using re-

source selection functions. Trends in Ecology and Evolution 14:268–272.

Brandes, U. and T. Erlebach, 2005. Network analysis: methodological foundations.

Springer, Berlin Heidelberg, Germany.

Breckling, B., U. Middelhoff, and H. Reuter, 2006. Individual-based models as tools

for ecological theory and application: understanding the emergence of organisational

properties in ecological systems. Ecological Modelling 194:102–113.

Brinkmeier, M. and T. Schank, 2005. Network statistics. In U. Brandes and T. Er-

lebach, editors, Network Analysis: Methodological Foundations, pages 293–316.

Springer-Verlag, Berlin Heidelberg, Germany.

Brooks, C. P., 2003. A scalar analysis of landscape connectivity. Oikos 102:433–439.

Brooks, C. P., 2006. Quantifying population substructure: extending the graph-

theoretic approach. Ecology 87:864–872. BIBLIOGRAPHY 141

Brooks, C. P., A. Antonovics, and T. Keitt, 2008. Spatial and temporal heterogene-

ity explain disease dynamics in a spatially explicit network model. The American

Naturalist 172:149–159.

Brudvig, L. A., E. I. Damschen, J. J. Tewksbury, N. M. Haddad, and D. J. Levey,

2009. Landscape connectivity promotes plant biodiversity spillover into non-target

habitats. Proceedings of the National Academy of Sciences of the United States of

America 106:9328–9332.

Bunn, A. G., D. L. Urban, and T. H. Keitt, 2000. Landscape connectivity: a conserva-

tion application of graph theory. Journal of Environmental Management 59:265–278.

Buskirk, S. W. and L. L. McDonald, 1989. Analysis of variability in home-range size of

the American marten. Journal of Wildlife Management 53:997–1004.

Buskirk, S. W. and R. A. Powell, 1994. Habitat ecology of fishers and American

martens. In S. Buskirk, A. Harestad, M. Raphael, and R. Powell, editors, Martens,

sables, and fishers biology and conservation, pages 283–296. Cornell University Press,

Ithaca.

Buskirk, S. W. and L. F. Ruggiero, 1994. Marten. In L. Ruggiero, K. Aubry,

S. Buskirk, L. Jack Lyon, and W. Zielinski, editors, The scientific basis for conserv-

ing forest carnivores. General Technical Report, pages 7–37. RM-254. Ft. Collins,

CO: USDA, Forest Service, Rocky Mountain Forest and Range Experiment Station.

Cabeza, M., 2003. Habitat loss and connectivity of reserve networks in probability

approaches to reserve design. Ecology Letters 6:665–672.

Cabeza, M., M. B. Araujo, R. J. Wilson, C. D. Thomas, M. J. R. Cowley, and

A. Moilanen, 2004. Combining probabilities of occurrence with spatial reserve de-

sign. Journal of Applied Ecology 41:252–262. BIBLIOGRAPHY 142

Cabeza, M. and A. Moilanen, 2001. Design of reserve networks and the persistence of

biodiversity. Trends in Ecology and Evolution pages 242–248.

Cabeza, M. and A. Moilanen, 2003. Site-Selection Algorithms and Habitat Loss. Con-

servation Biology 17:1402–1413.

Cabeza, M. and A. Moilanen, 2006. Replacement cost: a practical measure of site value

for cost-effective reserve planning. Biological Conservation 132:336–342.

Cain, D. H., K. Riiters, and K. Orvis, 1997. A multi-scale analysis of landscape statis-

tics. Landscape Ecology 122:199–212.

Calabrese, J. M. and W. F. Fagan, 2004. A comparison-shopper’s guide to connectivity

metrics. Frontiers in Ecology and the Environment 2:529–536.

Callaway, D. S., M. E. J. Newman, S. H. Strogatz, and D. J. Watts, 2000. Network

robustness and fragility: percolation on random graphs. Physical Review Letters

89:258–270.

Cantwell, M. D. and R. T. T. Forman, 1993. Landscape graphs - Ecological modeling

with graph-theory to detect configurations common to diverse landscapes. Landscape

Ecology 8:239–255.

Carignan, V. and M.-A. Villard, 2002. Selecting indicator species to monitor ecological

integrity: a review. Environmental Monitoring and Assessment 78:45–61.

Carpenter, S. R., T. M. Frost, A. R. Ives, J. F. Kitchell, and T. M. Kratz, 1994. Com-

plexity, cascades, and compensation in ecosystems, Global Environmental Forum. In

M. Yasuno and M. Watanabe, editors, Biodiversity: its complexity and role. National

Institute for Environmental Studies, Tokyo.

Carroll, C. and D. G. Miquelle, 2006. Spatial viability analysis of Amur tiger Panthera BIBLIOGRAPHY 143

tigris altaica in the Russian Far East: the role of protected areas and landscape ma-

trix in population persistence. Journal of Applied Ecology 43:1056–1068.

Carroll, C., R. F. Noss, P. C. Paquet, and N. H. Schumaker, 2003. Use of population

viability analysis and reserve selection algorithms in regional conservation plans.

Ecological Applications pages 1773–1789.

Chan, K. M. A., M. R. Shaw, D. R. Cameron, E. C. Underwood, and G. C. Daily,

2006. Conservation planning for ecosystem services. PLOS Biology 4:2138–2152.

Chapin, T. G., D. J. Harrison, and D. D. Katnik, 1998. Influence of landscape pattern

on habitat use by American marten in an industrial forest. Conservation Biology

12:1327–1337.

Chardon, J. P., F. Adriaensen, and E. Matthysen, 2003. Incorporating landscape el-

ements into a connectivity measure: a case study for the Speckled wood butterfly

(Pararge aegeria L.). Landscape Ecology 18:561–573.

Chetkiewicz, C.-L. B., C. Cassady St. Clair, and M. Boyce, 2006. Corridors for conser-

vation: integrating pattern and process. Annual Review of Ecological and Evolution-

ary Systems 37:317–342.

Clevenger, A. P., J. Wierzchowski, B. Chruszcz, and K. Gunson, 2002. GIS-generated,

expert-based models for identifying wildlife habitat linkages and planning mitigation

passages. Conservation Biology 16:503–514.

Costello, C. and S. Polasky, 2004. Dynamic reserve site selection. Resource and Energy

Economics 26:157–174.

Couvet, D., 2002. Deleterious effects of restricted gene flow in fragmented populations.

Conservation Biology 16:369–376. BIBLIOGRAPHY 144

Crooks, K. R. and M. Sanjayan, editors, 2006. Connectivity Conservation. Cambridge

University Press, Cambridge, UK.

Csuti, B., S. Polasky, P. H. Williams, R. L. Pressey, J. D. Camm, M. Kershaw, A. R.

Kiester, B. Downs, R. Hamilton, M. Huso, and K. Sahr, 1997. A comparison of re-

serve selection algorithms using data on terrestrial vertebrates in Oregon. Biological

Conservation 80:83–97.

Cuddington, K. M. and P. Yodzis, 2000. Diffusion-limited predatorprey dynamics in

Euclidean environments: an allometric individual-based model. Theoretical Popula-

tion Biology 58:259–278.

Cumming, S. G., P. J. Burton, and B. Klinkenberg, 1996. Boreal mixedwood forests

may have no “representative” areas: Some implications for reserve design. Ecography

19:162–180.

Cushman, S. A., K. McGarigal, and M. C. Neel, 2008. Parsimony in landscape metrics:

strength, universailty, and consistency. Ecological Indicators 8:691–703.

Cushman, S. A., K. S. McKelvey, J. Hayden, and M. K. Schwartz, 2006. Gene flow in

complex landscapes: Testing multiple hypotheses with causal modeling. American

Naturalist 168:486–499.

Damschen, E. I., L. A. Brudvig, N. M. Haddad, D. J. Levey, J. L. Orrock, and J. J.

Tewksbury, 2008. The movement ecology and dynamics of plant communities in frag-

mented landscapes. Proceedings of the National Academy of Sciences of the United

States of America 105:19078–19083.

Damschen, E. I., N. M. Haddad, J. L. Orrock, J. J. Tewksbury, and D. J. Levey, 2006.

Corridors increase plant species richness at large scales. Science 313:1284–1286. BIBLIOGRAPHY 145 de Roos, A. M., E. McCauley, and W. G. Wilson, 1998. Pattern formation and the

spatial scale of interaction between predator and their prey. Theoretical Population

Biology 53:108–130.

D’Eon, R. G., S. M. Glenn, I. Parfitt, and M.-J. Fortin, 2002. Landscape connectivity

as a function of scale and organism vagility in a real forested landscape. Conserva-

tion Ecology 6.

Didion, M., M.-J. Fortin, and A. Fall, 2007. Forest age structure as indicator of bo-

real forest sustainability under alternative management and fire regimes: a landscape

level sensitivity analysis. Ecological Modelling 200:45–58.

Diestel, R., 2006. Graph Theory. Graduate Texts in Mathematics. Birkhauser, third

edition.

Done, T. J. and R. E. Reichelt, 1998. Integrated coastal zone and fisheries ecosys-

tem management: Generic goals and performance indices. Ecological Applications

8:S110–S118.

Dorf, R. C. and J. A. Svoboda, 2003. Introduction to electric circuits. John Wiley and

Sons, New York, NY, 6th edition.

Drechsler, M., 2005. Probabilistic approaches to scheduling reserve selection. Biological

Conservation 122:253–262.

Drielsma, M., S. Ferrier, and G. Manion, 2007. A raster-based technique for analysing

habitat configuration: The cost-benefit approach. Ecological Modelling 202:324–332.

Driezen, K., F. Adriaensen, C. Rondinini, C. P. Doncaster, and E. Matthysen, 2007.

Evaluating least-cost model predictions with empirical dispersal data: A case-study

using radiotracking data of hedgehogs (Erinaceus europaeus). Ecological Modelling

209:314–322. BIBLIOGRAPHY 146

Dunne, J. A., R. J. Williams, and N. D. Martinez, 2002. Network structure and bio-

diversity loss in food webs: robustness increases with connectance. Ecology Letters

5:558–567.

Early, R. and C. D. Thomas, 2007. Multispecies conservation planning: identifying

landscapes for the conservation of viable populations using local and continental

species priorities. Journal of Applied Ecology 44:253–262.

Economo, E. P. and T. H. Keitt, 2008. Species diversity in neutral metacommunities: a

network approach. Ecology Letters 11:52–62.

Epps, C. W., J. D. Wehausen, V. C. Bleich, S. G. Torres, and J. S. Brashares, 2007.

Optimizing dispersal and corridor models using landscape genetics. Journal of Ap-

plied Ecology 44:714–724.

Estrada-Pena, A., 2002. Understanding the relationships between landscape connec-

tivity and abundance of Ixodes ricinus ticks. Experimental and Applied Acarology

28:239–248.

Estrada-Pena, A., J. Quilez, E. del Cacho, and C. S. Acedo, 2006. A framework to map

abundance of tick metapopulations. Ecological Modelling 193:663–674.

Fagan, W. F., 2002. Connectivity, fragmentation, and extinction risk in dendritic

metapopulations. Ecology 83:3243–3249.

Fagan, W. F. and J. M. Calabrese, 2006. Quantifying connectivity: balancing metric

performance with data requirements. In K. R. Crooks and M. Sanjayan, editors,

Connectivity conservation, pages 297–317. Cambridge University Press, Cambridge.

Fagan, W. F., R. S. Cantrell, and C. Cosner, 1999. How habitat edges change species

interactions. The American Naturalist 153:165–182. BIBLIOGRAPHY 147

Fahrig, L., 1997. Relative effects of habitat loss and fragmentation on population ex-

tinction. The Journal of Wildlife Management 61:603–610.

Fahrig, L., 1998. When does fragmentation of breeding habitat affect population sur-

vival? Ecological Modelling 105:273–292.

Fahrig, L., 2003. Effects of habitat fragmentation on biodiversity. Annual Review of

Ecological and Evolutionary Systems 34:487–515.

Fahrig, L., 2007. Non-optimal animal movement in human-altered landscapes. Func-

tional Ecology 21:1003–1015.

Fall, A. and J. Fall, 2001. A domain-specific language for models of landscape dynam-

ics. Ecological Modelling 141:1–18.

Fall, A., M.-J. Fortin, D. D. Kneeshaw, S. H. Yamasaki, C. Messier, L. Bouthillier, and

C. Smyth, 2004. Consequences of various landscape-scale ecosystem management

strategies and fire cycles on age-class structure and harvest in boreal forests. Cana-

dian Journal of Forest Research 34:310–322.

Fall, A., M.-J. Fortin, M. Manseau, and D. O’Brien, 2007. Spatial graphs: principles

and applications for habitat connectivity. Ecosystems 10:448–461.

FAO, 2006. Global Forest Resources Assessment 2005, main report. Progress Towards

Sustainable Forest Management. Technical report, FAO Forestry Paper 147.

Farmer, A. H. and J. A. Wiens, 1998. Optimal migration schedules depend on the

landscape and the physical environment: a dynamic modeling view. Journal of Avian

Biology 29:405–415.

Ferrari, J. R. and T. R. Lookingbill, 2009. Initial conditions and their effect on inva-

sion velocity across heterogeneous landscapes. Biological Invasions 11:1247–1258. BIBLIOGRAPHY 148

Ferrari, J. R., T. R. Lookingbill, and M. C. Neel, 2007. Two measures of landscape-

graph connectivity: assessment across gradients in area and configuration. Landscape

Ecology 22:1315–1323.

Ferrari, M. J., S. Bansal, L. A. Meyers, and O. N. Bjornstad, 2006. Network fragility

and the geometry of herd immunity. Proceedings of the Royal Society B-Biological

Sciences 273:2743–2748.

Ferreras, P., 2001. Landscape structure and asymmetrical inter-patch connectivity in

a metapopulation of the endangered Iberian lynx. Biological Conservation 100:125–

136.

Fischer, J. and D. B. Lindenmayer, 2007. Landscape modification and habitat fragmen-

tation: a synthesis. Global Ecology and Biogeography 16:265–280.

Foley, J. A., R. DeFries, G. P. Asner, C. Barford, G. Bonan, S. R. Carpenter, F. S.

Chapin, M. T. Coe, G. C. Daily, H. Gibbs, J. H. Helkowski, T. Holloway, E. A.

Howard, C. J. Kucharik, C. Monfreda, J. A. Patz, I. C. Prentice, N. Ramankutty,

and P. K. Snyder, 2005. Global consequences of land use. Science 309:570–574.

Folke, C., 2003. Reserves and resilience - from single equilibrium to complex systems.

Ambio 32:379–379.

Folke, C., S. Carpenter, B. Walker, M. Scheffer, T. Elmqvist, L. Gunderson, and C. S.

Holling, 2004. Regime shifts, resilience, and biodiversity in ecosystem management.

Annual Review of Ecology, Evolution, and Systematics 35:557–581.

Forman, R. T. T., 1995. Land Mosaics: The Ecology of Landscapes and Regions. Cam-

bridge University Press, Cambridge.

Forman, R. T. T. and M. Gordon, 1986. Landscape Ecology. John Wiley and Sons,

Inc, New York. BIBLIOGRAPHY 149

Fortuna, M. A., C. Gomez-Rodriguez, and J. Bascompte, 2006. Spatial network struc-

ture and amphibian persistence in stochastic environments. Proceedings of the Royal

Society B-Biological Sciences 273:1429–1434.

Franc, A., 2004. Metapopulation dynamics as a contact process on a graph. Ecological

Complexity 1:49–63.

Gaston, K. J., K. Charman, S. F. Jackson, P. R. Armsworth, A. Bonn, R. A. Briers,

C. S. Q. Callaghan, R. Catchpole, J. Hopkins, W. E. Kunin, J. Latham, P. Opdam,

R. Stoneman, D. A. Stroud, and R. Tratt, 2006. The ecological effectiveness of pro-

tected areas: The United Kingdom. Biological Conservation 132:76–87.

Gaston, K. J., R. L. Pressey, and C. R. Margules, 2002. Persistence and vulnerability:

retaining biodiversity in the landscape and in protected areas. Journal of Biosciences

27:361–384.

Gerber, L. R., L. W. Botsford, A. Hastings, H. P. Possingham, S. D. Gaines, S. R.

Palumbi, and S. Andelman, 2003. Population models for marine reserve design: a

retrospective and prospective synthesis. Ecological Applications 13:S47–S64.

Gilpin, M. E. and M. E. Soul´e,1986. Minimum viable populations: processes of species

extinction. In M. Soul´e,editor, Conservation biology: the science of scarcity and

diversity, pages 19–34. Sinauer Associates Inc., Sunderland, Massachusetts.

Gonzales, E. K. and S. E. Gergel, 2007. Testing assumptions of cost surface analysis -

a tool for invasive species management. Landscape Ecology 22:1155–1168.

Gonzalez, A. and E. Chaneton, 2002. Heterotrophic extinction, abundance and produc-

tivity in a fragmented microecosystem. Journal of Animal Ecology 71:594–603.

Goodwin, B. J. and L. Fahrig, 2002a. Effect of landscape structure on the movement BIBLIOGRAPHY 150

behaviour of a specialized goldenrod beetle, Trirhabda borealis. Canadian Journal of

Zoology 80:24–35.

Goodwin, B. J. and L. Fahrig, 2002b. How does landscape structure influence land-

scape connectivity? Oikos 99:552–570.

Graham, C. H., 2001. Factors influencing movement patterns of Keel-billed Toucans in

a fragmented tropical landscape in southern Mexico. Conservation Biology 15:1789–

1798.

Green, D. G. and S. Sadedin, 2005. Interactions matter - complexity in landscapes and

ecosystems. Ecological Complexity 2:117–130.

Groeneveld, R., 2005. Economic considerations in the optimal size and number of re-

serve sites. Ecological Economics 52:219–228.

Gross, J. L. and J. Yellen, 2004. Handbook of graph theory. CRC press.

Guichard, F., S. A. Levin, A. Hastings, and D. Siegel, 2004. Toward a dynamic meta-

community approach to marine reserve theory. Bioscience 54:1003–1011.

Guilhaumon, F., O. Gimenez, K. J. Gaston, and D. Mouillot, 2008. Taxonomic and

regional uncertainty in species-area relationships and the identification of richness

hotspots. Proceedings of the National Academy of Sciences of the United States of

America 105:15458–15463.

Gunderson, L. H., 2000. Ecological resilience - in theory and application. Annual Re-

view of Ecology and Evolutionary Systems 31:425–439.

Gustafson, E. J., 1998. Clustering timber harvests and the effect of dynamic forest

management policy on forest fragmentation. Ecosystems 1:484–492.

Haddad, N. M. and J. J. Tewksbury, 2005. Low quality habitat corridors as movement

conduits for two butterfly species. Ecological Applications 15:250–257. BIBLIOGRAPHY 151

Hall, L. S., P. Krausman, and M. L. Morrison, 1997. The habitat concept and a plea

for standard terminology. Wildlife Society Bulletin 25:173–182.

Halpin, P. N. and A. G. Bunn, 2000. Using GIS to compute a least-cost distance ma-

trix: A comparison of terrestrial and marine ecological applications. In Proceedings

of the 2000 ESRI User Conference, pages 1–19.

Hanski, I., 1998a. Connecting the parameters of local extinction and metapopulation

dynamics. Oikos 83:390–396.

Hanski, I., 1998b. Metapopulation dynamics. Nature 396:4149.

Hanski, I., 1999a. Habitat connectivity, habitat continuity, and metapopulations in

dynamic landscapes. Oikos 87:209–219.

Hanski, I., 1999b. Metapopulation Ecology. Oxford University Press, New York.

Hanski, I. and O. E. Gaggiotti, 2004. Metapopulation biology: past, present, and fu-

ture. In I. Hanski and O. Gaggiotti, editors, Ecology, genetics, and evolution of

metapopulations, pages 3–22. Elsevier-Academic Press, San Diego.

Hanski, I. and O. Ovaskainen, 2000. The metapopulation capacity of a fragmented

landscape. Nature 404:755–758.

Harary, F., 1969. Graph theory. Addison-Wesley series in mathematics. Addison-

Wesley Reading, Massachusetts.

H´ely, C., Y. Bergeron, and M. D. Flannigan, 2000. Effects of stand composition on

fire hazard in mixed-wood Canadian boreal forest. Journal of Vegetation Science

11:812–824.

Hilty, J. A., W. Z. J. Lidicker, and A. M. Merenlender, 2006. Corridor Ecology: The

Science and Practice of Linking Landscapes for Biodiversity Conservation. Island

Press, Washington, D.C. BIBLIOGRAPHY 152

Holland, M. D. and A. Hastings, 2008. Strong effect of dispersal network structure on

ecological dynamics. Nature 456:792–795.

Holling, C. S., 1973. Resilience and stability of ecological systems. Annual Review of

Ecology, Evolution, and Systematics 41:1–23.

Holt, R., 1997. From metapopulation dynamics to community sructure: some conse-

quences of spatial heterogeneity. In I. Hanski and M. Gilpin, editors, Metapopula-

tion Biology: Ecology, Genetics, and Evolution, pages 149–164. Academic Press, San

Diego.

Holzkamper, A., A. Lausch, and R. Seppelt, 2006. Optimizing landscape configuratlion

to enhance habitat suitability for species with contrasting habitat requirements. Eco-

logical Modelling 198:277–292.

Hoopes, M. F., R. Holt, and M. Holyoak, 2005. The effects of spatial processes on two

species interactions. In M. Holyoak, M. Leibold, and R. Holt, editors, Metacom-

munities: Spatial Dynamics and Ecological Comunities, pages 35–67. University of

Chicago Press, Chicago.

Hufkens, K., R. Ceulemans, and P. Scheunders, 2008. Estimating the ecotone width in

patchy ecotones using a sigmoid wave approach. Ecological Informatics 3:97–104.

Hunter, J., M. L., 1991. Coping with ignorance: the coarse-filter strategy for maintain-

ing biological diversity. In K. Kohm, editor, Balancing on the brink of extinction,

pages 266–281. Island Press, Washington, D.C.

Ims, R. A., 1995. Movement patterns related to spatial structures. In L. Hansson,

L. Fahrig, and G. Merriam, editors, Mosaic landscapes and ecological processes, pages

85–109. Chapman, London. BIBLIOGRAPHY 153

IUCN, 1994. Guidelines for Protected Area Management Categories. Technical report,

CNPPA with the assistance of WCMC.

James, P. M. A., M.-J. Fortin, A. Fall, D. Kneeshaw, and C. Messier, 2007. The effects

of spatial legacies following shifting management practices and fire on boreal forest

age structure. Ecosystems 10:1261–1277.

James, P. M. A., B. Rayfield, M.-J. Fortin, and A. Fall, 2005. Reserve network de-

sign combining spatial graph theory and species’ spatial requirements. Geomatica

59:323–333.

Janssen, M. A., O. Bodin, J. M. Anderies, T. Elmqvist, H. Ernston, R. R. J. McAl-

lister, P. Olsson, and P. Ryan, 2006. Toward a network perspective of the study of

resilience in social-ecological systems. Ecology and Society 11.

Jeong, H., S. P. Mason, A.-L. Barab´asi,and Z. N. Oltvai, 2001. Lethality and central-

ity in networks. Nature 411:41.

Johnson, C. J. and M. P. Gillingham, 2004. Mapping uncertainty: sensitivity of wildlife

habitat ratings to expert opinion. Journal of Applied Ecology 41:1032–1041.

Johnson, E. A., 1992. Fire and vegetation dynamics: studies from the North American

Boreal forest. Cambridge University Press, Cambridge.

Johnson, E. A., K. Miyanishi, and S. R. J. Bridge, 2001. Wildfire Regime in the Boreal

Forest and the Idea of Suppression and Fuel Buildup. Conservation Biology 15:1554–

1557.

Johnson, E. A. and C. E. Van Wagner, 1985. The theory and use of two fire history

models. Canadian Journal of Forest Research 15:214–240.

Jord´an,F., 2000. A reliability-theory approach to corridor design. Ecological Modelling

128:211–220. BIBLIOGRAPHY 154

Jord´an,F., 2001. Adding function to structure - comments on Palmarola landscape

connectivity. Community Ecology 2:133–135.

Jord´an,F., 2003. Quantifying landscape connectivity: key patches and key corridors.

Ecosystems and Sustainable Development IV 18-19:883–891.

Jord´an,F., T. Magura, B. Tothmeresz, V. Vasas, and V. Kodobocz, 2007. Carabids

(Coleoptera : Carabidae) in a forest patchwork: a connectivity analysis of the Bereg

Plain landscape graph. Landscape Ecology 22:1527–1539.

Kareiva, P. and U. Wennergren, 1995. Connecting landscape patterns to ecosystem and

population processes. Nature 373:299–302.

Kautz, R., R. Kawula, T. Hoctor, J. Comiskey, D. Jansen, D. Jennings, J. Kasbohm,

F. Mazzotti, R. McBride, L. Richardson, and K. Root, 2006. How much is enough?

Landscape-scale conservation for the Florida panther. Biological Conservation

130:118–133.

Keil, J. M. and C. A. Gutwin, 1992. Classes of graphs which approximate the complete

Euclidean graph. Discrete Computational Geometry 7:13–28.

Keitt, T., D. Urban, and B. T. Milne, 1997. Detecting critical scales in fragmented

landscapes. Ecology and Society 1:4.

Keitt, T. H., 2003. Network Theory: an evolving approach to landscape conservation.

In V. Dale, editor, Ecological Modeling for Resource Management, pages 125–134.

Springer-Verlag, New York, NY.

Kim, S., W. Noh, and S. An, 2003. Multi-path ad hoc routing considering path redun-

dancy. In I. C. Society, editor, IEEE Inetrnational Symposium on Computers and

Communication (ISCC’03), pages 45–50. BIBLIOGRAPHY 155

Klug, J. L., J. M. Fischer, A. R. Ives, and B. Dennis, 2000. Compensatory dynamics in

planktonic community responses to pH perturbations. Ecology 81:387–398.

Knappen, J. P., M. Scheffer, and B. Harms, 1992. Estimating habitat isolation in land-

scape planning. Landscape and Urban Planning 23:1–16.

Koelle, K. and J. Vandermeer, 2005. Dispersal-induced desynchronization: from

metapopulations to metacommunities. Ecology Letters 8:167–175.

Koenig, W. D., D. V. Vuren, and P. N. Hooge, 1996. Detectability, philoparty, and the

distribution of dispersal distances in vertebrates. Trends in Ecology and Evolution

11:514–517.

Kotliar, N. B. and J. A. Wiens, 1990. Multiple scales of patchiness and patch struc-

ture: a hierarchical framework for the study of heterogeneity. Oikos 59:253–260.

Krebs, C. J., 2002. Two complementary paradigms for analysing population dynam-

ics. Philosophical Transactions of the Royal society of London, Series B - Biological

Sciences 357:1211–1219.

Kremen, C., A. Cameron, A. J. Moilanen, S. Philips, C. D. Thomas, H. Beentje,

J. Dransfeld, B. L. Fisher, F. Glaw, T. Good, G. Harper, R. J. Hijmans, D. C. Lees,

E. Louis Jr., R. A. Nussbaum, C. Raxworthy, A. Razafimpahanana, G. Schatz,

M. Vences, D. R. Vieites, P. C. Wright, and M. L. Zjhra, 2008. Aligning conserva-

tion priorities across taxa in Madagascar, a biodiversity hotspot, with high-resolution

planning tools. Science 320:222–226.

Landres, P. B., J. Verner, and J. W. Thomas, 1988. Ecological uses of vertebrate indi-

cator species: a critique. Conservaion Biology 2:316–328.

LaRue, M. A. and C. K. Nielsen, 2008. Modelling potential dispersal corridors for BIBLIOGRAPHY 156

cougars in midwestern North America using least-cost path methods. Ecological

Modelling 212:372–381.

Lees, A. C. and C. A. Peres, 2008. Conservation value of remnant riparian forest cor-

ridors of varying quality for Amazonian birds and mammals. Conservation Biology

22:439–449.

Lefort, P. S., S. Gauthier, and Y. Bergeron, 2003. The influence of fire weather and

land use on the fire activity of the lake Abitibi area, Eastern Canada. Forest Science

49:509–521.

Leroux, S. J., F. K. A. Schmiegelow, S. G. Cumming, R. B. Lessard, and J. Nagy,

2007a. Accounting for system dynamics in reserve design. Ecological Applications

17:1954–1966.

Leroux, S. J., K. A. Schmiegelow, R. B. Lessard, and S. G. Cumming, 2007b. Min-

imum dynamic reserves: a framework for determining reserve size in ecosystems

structured by large disturbances. Biological Conservation 138:464–473.

Levins, R., 1969. Some demographic and genetic consequences of environmental het-

erogeneity for biological control. Bulletin of the Entomological Society of America

15:237–240.

Li, H. and J. F. Reynolds, 1994. A simulation experiment to quantify spatial hetero-

geneity in categorical maps. Ecology 75:2446–2455.

Lima, S. L. and P. A. Zollner, 1996. Towards a behavioral ecology of ecological land-

scapes. Trends in Ecology and Evolution 11:131–136.

Lindenmayer, D. B. and J. F. Franklin, 2002. Conserving Forest Biodiversity. Island

Press, Washington, D.C. BIBLIOGRAPHY 157

Lindenmayer, D. B., J. F. Franklin, and J. Fischer, 2006. General management princi-

ples and a checklist of strategies to guide forest biodiversity conservation. Biological

Conservation 131:433–445.

MacArthur, R. H. and E. O. Wilson, 1967. The Theory of Island Biogeography. Prince-

ton University Press, Princeton, New Jersey.

Magle, S. B., D. M. Theobald, and K. R. Crooks, 2009. A comparison of metrics pre-

dicting landscape connectivity for a highly interactive species along an urban gradi-

ent in Colorado, USA. Landscape Ecology 24:267–280.

Mangan, S. and U. Alon, 2003. Structure and function of the feed-forward loop net-

work motif. Proceedings of the National Academy of Sciences 100:11980–11985.

Manly, B. F. J., L. L. McDonald, D. L. Thomas, T. L. McDonald, and W. P. Erickson,

2002. Resource selection by animals: statistical design and analysis for field studies.

Kluwer Academic Publishers, Dordrecht, The Netherlands, second edition.

Manseau, M., A. Fall, D. OBrien, and M.-J. Fortin, 2002. National Parks and the pro-

tection of woodland caribou: a multi-scale landscape analysis method. Research

Links 10:24–28.

Marcot, B. G., 1998. Assessing the connectivity of habitat patches by matrix analysis

and graph theory. Analysis notes 8:12–16.

Marcot, B. G. and P. Z. Chinn, 1982. Use of graph theory measures for assessing diver-

sity of wildlife habitat. In R. Lamberson, editor, Mathematical Models of Renewable

Resources, Proceedings of the First Pacific Coast Conference on Mathematical Mod-

els of Renewable Resources. Humboldt State University, Arcata, CA.

Margules, C. R., A. O. Nicholls, and M. B. Usher, 1994. Apparent species turnover, BIBLIOGRAPHY 158

probability of extinction and the selection of nature-reserves - a case-study of the

Ingleborough Limestone Pavements. Conservation Biology 8:398–409.

Margules, C. R. and R. L. Pressey, 2000. Systematic conservation planning. Nature

405:243–253.

Matisziw, T. C. and A. T. Murray, 2009. Connectivity change in habitat networks.

Landscape Ecology 24:89–100.

Matlack, G. R. and J. Monde, 2004. Consequences of low mobililty in spatially and

temporally heterogeneous ecosystems. Journal of Ecology 92:1025–1035.

Matter, S. F., 2003. Modeling the densityarea relationship in a dynamic landscape:

an examination for the beetle Tetraopes tetraophthalmus and a generalized model.

Ecological Modelling 169:103–117.

May, R., 1994. The effects of spatial scale on ecological questions and answers. In

P. Edwards, R. May, and N. Webb, editors, Large-scale Ecology and Conservation

Biology, pages 1–17. Blackwell Scientific Publications, Oxford.

May, R. M., 2006. Network structure and the biology of populations. Trends in Ecology

and Evolution 21:394–399.

McCann, K., A. Hastings, and G. Huxel, 1998. Weak trophic interactions and the bal-

ance of nature. Nature 395:794–798.

McDonnell, M. D., H. P. Possingham, I. R. Ball, and E. A. Cousins, 2002. Mathemat-

ical methods for spatially cohesive reserve design. Environmental Modeling and As-

sessment 7:107–114.

McGarigal, K. and S. Cushman, 2002. Comparative evaluation of experimental ap-

proaches to the study of habitat fragmentation effects. Ecological Applications

12:335–345. BIBLIOGRAPHY 159

McIntyre, N. E. and J. A. Wiens, 1999. Interactions between landscape structure and

animal behavior: the roles of heterogeneously distributed resources and food depriva-

tion on movement patterns. Landscape Ecology 14:437–447.

McLaren, M. A., I. D. Thompson, and J. A. Baker, 1998. Selection of vertebrate

wildlife indicators for monitoring sustainable forest management in Ontario. Forestry

Chronicle 74:241–248.

McNeely, J. A., editor, 1993. Parks for Life: Report of the IVth World Congress on

National Parks and Protected Areas. IUCN Communications Division, Gland,

Switzerland.

McRae, B. H., 2006. Isolation by resistance. Evolution 60:1551–1561.

McRae, B. H. and P. Beier, 2007. Circuit theory predicts gene flow in plant and anmial

populations. Proceedings of the National Academy of Sciences 104:19885–19890.

McRae, B. H., B. G. Dickson, T. H. Keitt, and V. B. Shah, 2008. Using circuit theory

to model connectivity in ecology, evolution, and conservation. Ecology 89:2712–2724.

Meir, E., S. Andelman, and H. P. Possingham, 2004. Does conservation planning mat-

ter in a dynamic and uncertain world? Ecology Letters 7:615–622.

Meli´an,C. J. and J. Bascompte, 2002. Complex networks: two ways to be robust?

Ecology Letters 5:707–708.

M˝onkk˝onen,M. and P. Reunanen, 1999. On critical thresholds in landscape connectiv-

ity: a management perspective. Oikos 84:302–305.

Micheli, F., P. Amarasekare, J. Bascompte, and L. Gerber, 2004. Including species

interactions in the design and evaluation of marine reserves: some insights from a

predator-prey model. Bulletin of Marine Science 74:653–669. BIBLIOGRAPHY 160

Miller, M. S. and R. E. Russell, 2004. Sharp edges, matrix permeability, and individual

behavior: considerations for improving estimates of landscape connectivity. In R. S.

Moore and J. E., editors, Conserving biodiversity in agricultural landscapes: model-

based planning tools for systems with sharp edges, pages 121–138. Purdue University

Press, West Lafayette.

Milo, R., S. Itzkovitz, N. Kashtan, R. Levitt, S. Shen-Orr, I. Ayzenshtat, M. Shef-

fer, and U. Alon, 2004. Superfamilies of evolved and designed networks. Science

303:1538–1542.

Milo, R., S. Shen-Orr, S. Itzkovitz, N. Kashtan, D. Chklovskii, and U. Alon, 2002. Net-

work motifs: simple building blocks of complex networks. Science 298:824827.

Minor, E. S. and D. L. Urban, 2007. Graph theory as a proxy for spatially explicit

population models in conservation planning. Ecological Applications 17:1771–1782.

Minor, E. S. and D. L. Urban, 2008. A graph-theory frarmework for evaluating land-

scape connectivity and conservation planning. Conservation Biology 22:297–307.

Moilanen, A., 2007. Landscape Zonation, benefit functions and target-based planning:

unifying reserve selection strategies. Biological Conservation 134:571–579.

Moilanen, A. and M. Cabeza, 2002. Single-species dynamic site selection. Ecological

Applications 12:913–926.

Moilanen, A., A. M. A. Franco, R. I. Early, R. Fox, B. Wintle, and C. D. Thomas,

2005. Prioritizing multiple-use landscapes for conservation: methods for large multi-

species planning problems. Proceedings of the Royal Society B-Biological Sciences

272:1885–1891.

Moilanen, A. and H. Kujala, 2006. Zonation spatial conservation planning framework

and software v. 1.0, User manual. Technical report. BIBLIOGRAPHY 161

Moilanen, A., J. Leathwick, and J. Elith, 2008. A method for freshwater conservation

prioritization. Freshwater Biology 53:577–592.

Moilanen, A. and M. Nieminen, 2002. Simple connectivity measures in spatial ecology.

Ecology 83:1131–1145.

Moilanen, A., M. C. Runge, J. Elith, A. Tyre, Y. Carmel, E. Fegraus, B. A. Wintle,

M. Burgman, and Y. Ben-Haim, 2006. Planning for robust reserve networks using

uncertainty analysis. Ecological Modelling 199:115–124.

Moilanen, A. and B. A. Wintle, 2006. Uncertainty analysis favours selection of spatially

aggregated reserve networks. Biological Conservation 129:427–434.

Moilanen, A. and B. A. Wintle, 2007. The boundary-quality penalty: a quantitative

method for approximating species responses to fragmentation in reserve selection.

Conservation Biology 21:355–364.

Moulton, V., C. Semple, and M. Steel, 2007. Optimizing phylogenetic diversity under

constraints. Journal of Theoretical Biology 246:186–194.

Mueller, T. and W. F. Fagan, 2008. Search and navigation in dynamic environments -

from individual behaviours to population distributions. Oikos 117:654–664.

Newman, M., A.-L. Barab´asi,and D. Watts, editors, 2006. The structure and dynamics

of networks. Princeton University Press, Princeton, NJ.

Newman, M. E. J., 2003. The structure and function of complex networks. SIAM

Review 45:167–256.

Noss, R. F., 1987. From plant communities to landscapes in conservation inventories: a

look at the Nature Conservancy (USA). Biological Conservation 41:11–37. BIBLIOGRAPHY 162

O’Brien, D., M. Manseau, A. Fall, and M.-J. Fortin, 2006. Testing the importance

of spatial configuration of winter habitat for woodland caribou: An application of

graph theory. Biological Conservation 130:70–83.

Oetting, J. B., A. L. Knight, and G. R. Knight, 2006. Systematic reserve design as a

dynamic process: F-TRAC and the Florida Forever program. Biological Conserva-

tion 128:37–46.

Oksanen, L., 2001. Logic of experiments in ecology: is pseudoreplication a pseudoissue.

Oikos 94:27–38.

Olden, J. D. and N. L. Poff, 2003. Toward a mechanistic understanding and prediction

of biotic homogenization. The American Naturalist 162:442–460.

Olson, D. M. and E. Dinerstein, 1998. The global 200: A representation approach to

conserving the Earth’s most biologically valuable ecoregions. Conservation Biology

12:502–515.

Palmer, M. A., D. A. Falk, and J. B. Zedler, 2006. Ecological theory and restoration

ecology. In D. Falk, M. Palmer, and J. B. Zedler, editors, Foundations of Restoration

Ecology, pages 1–10. Island Press, Washington D.C.

Pascual-Hortal, L. and S. Saura, 2006. Comparison and development of new graph-

based landscape connectivity indices: towards the priorization of habitat patches and

corridors for conservation. Landscape Ecology 21:959–967.

Pascual-Hortal, L. and S. Saura, 2007. Impact of spatial scale on the identification of

critical habitat patches for the maintenance of landscape connectivity. Landscape and

Urban Planning 83:176–186.

Pascual-Hortal, L. and S. Saura, 2008. Integrating landscape connectivity in broad-

scale forest planning through a new graph-based habitat availability methodology: BIBLIOGRAPHY 163

application to capercaillie (Tetrao urogallus) in Catalonia (NE Spain). European

Journal of Forest Research 127:23–31.

Payer, D., D. J. Harrison, and D. Phillips, 2004. Territoriality and home-range fidelity

of american marten in relation to timber harvesting and trapping. In D. Payer,

D. J. Harrison, A. Fuller, and G. Proulx, editors, Martens and Fishers (Martes) in

Human-Altered Environments: An International Perspective, pages 99–114. Springer

Inc., New York.

Peterson, G. D., 2002. Estimating resilience across landscapes. Conservation Ecology

6:http://www.consecol.org/vol6/iss1/art17/.

Pickett, S. T. A. and J. N. Thompson, 1978. Patch Dynamics and Design of Nature

Reserves. Biological Conservation 13:27–37.

Pinto, N. and T. H. Keitt, 2009. Beyond the least-cost path: evaluating corridor redun-

dancy using a graph-theoretic approach. Landscape Ecology 24:253–266.

Polis, G. A., 1998. Stability is woven by complex webs. Nature 395:744–745.

Possingham, H. P., I. R. Ball, and S. Andelman, 2000. Mathematical methods for

identifying representative reserve networks. In S. Ferson and M. Burgman, editors,

Quantitative Methods for Conservation Biology, pages 291–305. Springer-Verlag, New

York.

Potvin, F., L. Belanger, and K. Lowell, 2000. Marten habitat selection in a clearcut

boreal landscape. Conservation Biology 14:844–857.

Prakash, S. and A. M. de Roos, 2002. Habitat destruction in a simple predator-prey

patch model: How predators enhance prey persistence and abundance. Theoretical

Population Biology 62:231–249.

Pressey, R. L., 1999. Systematic conservation planning for the real world. Parks 9:1–6. BIBLIOGRAPHY 164

Pressey, R. L., R. M. Cowling, and M. Rouget, 2003. Formulating conservation tar-

gets for biodiversity pattern and process in the Cape Floristic Region, South Africa.

Biological Conservation 112:99–127.

Pressey, R. L., H. P. Possingham, and C. R. Margules, 1996. Optimality in reserve

selection algorithms: when does it matter and how much? Biological Conservation

76:259–267.

Pressey, R. L. and K. H. Taffs, 2001. Sampling of land types by protected areas: Three

measures of effectiveness applied to western New South Wales. Biological Conserva-

tion pages 105–117.

Proulx, S. R., D. E. L. Promislow, and P. C. Phillips, 2005. Network thinking in ecol-

ogy and evolution. Trends in Ecology and Evolution 20:345–353.

Pulliam, H. R., 1988. Sources, sinks, and population regulation. American Naturalist

132:652–661.

Pulliam, H. R. and B. R. Johnson, 2002. Ecology’s New Paradigm: What Does It Offer

Designers and Planners? In B. Johnson and K. Hill, editors, Ecology and Design,

Frameworks for Learning. Island Press, Washington, D.C.

Pyke, C. R., S. J. Andelman, and G. Midgley, 2005. Identifying priority areas for bio-

climatic representation under climate change: a case study for Proteaceae in the

Cape Floristic Region, South Africa. Biological Conservation 125:1–9.

Rayfield, B., M.-J. Fortin, and A. Fall, accepted. The sensitivity of least-cost habitat

graphs to relative cost surface values. Landscape Ecology .

Rayfield, B., P. M. A. James, A. Fall, and M.-J. Fortin, 2008. Comparing static ver-

sus dynamic protected areas in the Qu´ebec boreal forest. Biological Conservation

141:438–449. BIBLIOGRAPHY 165

Rayfield, B., A. Moilanen, and M.-J. Fortin, 2009. Incorporating consumer-resource

spatial interactions in reserve design. Ecological Modelling 220:725–733.

Rhodes, M., G. W. Wardell-Johnson, M. P. Rhodes, and B. Raymond, 2006. Applying

network analysis to the conservation of habitat trees in urban environments: a case

study from Brisbane, Australia. Conservation Biology 20:861–870.

Ricketts, T. H., 2001. The matrix matters: effective isolation in fragmented landscapes.

American Naturalist 158:87–99.

Ricotta, C., A. Stanisci, G. Avena, and C. Blasi, 2000. Quantifying the network con-

nectivity of landscape mosaics: a graph-theoretical approach. Community Ecology

1:89–94.

Riiters, K. H., R. V. ONeill, C. T. Hunsacker, J. D. Wickham, D. H. Yankee, S. P.

Timmins, K. B. Jones, and B. L. Jackson, 1995. A factor analysis of landscape pat-

tern and structure metrics. Landscape Ecology 10:23–39.

Riitters, K., J. Wickham, R. O’Neill, B. Jones, and E. Smith, 2000. Global-scale

patterns of forest fragmentation. Conservation Ecology 4:3. [online] URL:

http://www.consecol.org/vol4/iss2/art3/.

Roberge, J.-M. and P. Angelstam, 2004. Usefulness of the umbrella species concept as

a conservation tool. Conservaion Biology 18:76–85.

Rodrigues, A. S. L., R. D. Gregory, and K. J. Gaston, 2000. Robustness of reserve se-

lection procedures under temporal species turnover. Proceedings of the Royal Society

of London Series B-Biological Sciences 267:49–55.

Rothley, K., 2005. Finding and filling the ”cracks” in resis-

tance surfaces for least-cost modeling. Ecology and Society

10:http://www.ecologyandsociety.org/vol10/iss1/art4/. BIBLIOGRAPHY 166

Rouget, M., R. M. Cowling, A. T. Lombard, A. T. Knight, and G. I. H. Kerley, 2006.

Designing large-scale conservation corridors for pattern and process. Conservation

Biology 20:549–561.

Ruediger, W. C., 1985. Implementing a spotted owl management plan: the Gifford

Pinchot National Forest experience. In R. Gutierrez and A. Carey, editors, General

Technical Report PNW-185, pages 10–13. U.S. Department of Agriculture, Forest

Service, Pacific Northwest Forest and Range Experiment Station, Portland, Oregon.

Rush O’Hanley, J., R. L. Church, and J. Keith Gilless, 2007. Locating and protecting

critical reserve sites to minimize expected and worst-case losses. Biological Conserva-

tion 134:130–141.

Ryall, K. L. and L. Fahrig, 2006. Response of predators to loss and fragmentation of

prey habitat: a review of theory. Ecology Letters 87:1086–1093.

Sabo, J., 2008. Population viability and species interactions: life outside the single-

species vacuum. Biological Conservation 141.

Sadahiro, Y. and M. Umemura, 2001. A computational approach for the analysis of

changes in polygon statistics. Journal of Geographical Systems 3:137–154.

Saunders, D. A., R. J. Hobbs, and C. R. Margules, 1991. Biological consequences of

ecosystem fragmentation: a review. Conservation Biology 5:18–32.

Saura, S., 2008. Evaluating forest landscape connectivity through Conefor Sensinode

2.2. Patterns and Processes in Forest Landscapes: Multiple Use and Sustainable

Management pages 403–422.

Saura, S. and L. Pascual-Hortal, 2007. A new habitat availability index to integrate

connectivity in landscape conservation planning: comparison with existing indices

and application to a case study. Landscape and Urban Planning 83:91–103. BIBLIOGRAPHY 167

Saura, S. and J. Torne, 2009. Conefor Sensinode 2.2: A software package for quanti-

fying the importance of habitat patches for landscape connectivity. Environmental

Modelling & Software 24:135–139.

Schadt, S., F. Knauer, P. Kaczensky, E. Revilla, T. Wiegand, and L. Trepl, 2002. Rule-

based assessment of suitable habitat and patch connectivity for the Eurasian lynx.

Ecological Applications 12:1469–1483.

Schafer, C., 2001. Inter-reserve distance. Biological Conservation 100:215–227.

Schemske, D. W., B. C. Husband, M. H. Ruckelshaus, C. Goodwillie, I. M. Parker, and

J. G. Bishop, 1994. Evaluating approaches to the conservation of rare and endan-

gered plants. Ecology 75:584–606.

Schick, R. S. and S. T. Lindley, 2007. Directed connectivity among fish populations in

a riverine network. Journal of Applied Ecology 44:1116–1126.

Schindler, D., 1998. Replication versus realism: the need for ecosystem-scale experi-

ments. Ecosystems 1:323334.

Schultz, T. T. and L. A. Joyce, 1992. A spatial application of a marten habitat model.

Wildlife Society Bulletin 20:74–83.

Shafer, C. L., 2001. Inter-reserve distance. Biological Conservation 100:215–227.

Shen-Orr, S., R. Milo, S. Mangan, and U. Alon, 2002. Network motifs in the transcrip-

tional regulation network of Escherichia coli. Nature Genetics 31:64–68.

Sih, A., G. Englund, and D. Wooster, 1998. Emergent impacts of multiple predators on

prey. Trends in Ecology and Evolution 13:350–355.

Sinclair, A. R. E. and A. E. Byrom, 2006. Understanding ecosystem dynamics for con-

servation of biota. Journal of Animal Ecology 75:64–79. BIBLIOGRAPHY 168

Snyder, S. A., R. G. Haight, and C. S. ReVelle, 2004. Scenario optimization model for

dynamic reserve site selection. Environmental Modeling and Assessment 9:179–187.

Soul´e,M. E., J. A. Estes, J. Berger, and C. Martinez Del Rio, 2003. Ecological effec-

tiveness: conservation goals for interactive species. Conservation Biology 17:1238–

1250.

Soul´e,M. E., J. A. Estes, B. Miller, and D. L. Honnold, 2005. Strongly interacting

species. Conservation policy, management, and ethics. Bioscience 55:168–176.

Srinivasan, U. T., J. A. Dunne, J. Harte, and N. D. Martinez, 2007. Response of com-

plex food webs to realistic extinction sequences. Ecology 88:671–682.

Srivastava, D., J. Kolasa, J. Bengtsson, A. Gonzalez, S. Lawler, T. Miller, P. Munguia,

T. Romanuk, D. Schneider, and K. Trzcinski, 2004. Are natural microcosms useful

model systems for ecology? Trends in Ecology and Evolution 19:379–384.

Stevens, V. M., E. Polus, R. A. Wesselingh, N. Schtickzelle, and M. Baguette, 2004.

Quantifying functional connectivity: experimental evidence for patch-specific resis-

tance in the Natterjack toad (Bufo calamita). Landscape Ecology 19:829–842.

Stocks, B. J., J. A. Mason, J. B. Todd, E. M. Bosch, B. M. Wotton, B. D. Amiro,

M. D. Flannigan, K. G. Hirsch, K. A. Logan, D. L. Martell, and W. R. Skinner,

2002. Large forest fires in Canada, 1959 - 1997. Journal of Geophysical Research

108:8149.

Strange, N., B. Jellesmark Thorsen, and J. Bladt, 2006. Optimal reserve selection in a

dynamic world. Biological Conservation 131:33–41.

Strogatz, S. H., 2001. Exploring complex networks. Nature 410:268275.

Sutherland, G. D., D. T. OBrien, A. Fall, F. L. Waterhouse, A. S. Harestad, and J. B.

Buchanan, 2007. A framework to support landscape analyses of habitat supply and BIBLIOGRAPHY 169

effects on populations of forest-dwelling species: a case study based on the Northern

Spotted Owl. Technical report, British Columbia Ministry of Forests and Range

Technical Report 038. ISBN 978-0-7726-5677-3.

Svirezhev, Y. M., 2008. Nonlinearities in mathematical ecology: phenomena and mod-

els. Would we live in Volterra’s world? Ecological Modelling 216:89–101.

Swihart, R. K., Z. Feng, N. A. Slade, D. M. Mason, and T. M. Gehring, 2001. Effects

of habitat destruction and resource supplementation in a predator-prey metapopula-

tion model. Journal of Theoretical Biology 210:287–303.

Taylor, P. D., L. Fahrig, K. Henein, and G. Merriam, 1993. Connectivity is a vital ele-

ment of landscape structure. Oikos 68:571–573.

Tewksbury, J. J., D. J. Levey, N. M. Haddad, S. Sargent, J. L. Orrock, A. Weldon,

B. J. Danielson, J. Brinkerhoff, E. I. Damschen, and P. Townsend, 2002. Corridors

affect plants, animals, and their interactions in fragmented landscapes. Proceed-

ings of the National Academy of Sciences of the United States of America 99:12923–

12926.

Theobald, D. M., 2006. Exploring the functional connectivity of landscapes using land-

scape networks. In K. Crooks and M. Sanjayan, editors, Connectivity conservation,

pages 416–443. Cambridge University Press, Cambridge.

Thompson, I. D., 1991. Could marten become the Spotted Owl of Eastern Canada.

Forestry Chronicle 67:136–140.

Thompson, I. D., 1994. Marten populations in uncut and logged boreal forests in On-

tario. Journal of Wildlife Management 58:272–280.

Thompson, I. D., 2004. The importance of superior-quality wildlife habitats. The

Forestry Chronicle 80:75–81. BIBLIOGRAPHY 170

Thompson, I. D., 2006. Monitoring of biodiversity indicators in boreal forests: a need

for improved focus. Environmental Monitoring and Assessment 121:263273.

Thompson, I. D. and A. S. Harstad, 1994. Effects of logging on American martens and

models for habitat management. In S. Buskirk, A. Harestad, and M. Raphael, edi-

tors, Martens, sables, and fishers: biology and conservation, pages 355–367. Cornell

University Press, Ithaca.

Tischendorf, L., 1997. Modelling individual movements in heterogeneous landscapes:

potentials of a new approach. Ecological Modelling 103:33–42.

Tischendorf, L., 2001. Can landscape indices predict ecological processes consistently?

Landscape Ecology 16:235–254.

Tischendorf, L., D. J. Bender, and L. Fahrig, 2003. Evaluation of patch isolation met-

rics in mosaic landscapes for specialist vs. generalist dispersers. Landscape Ecology

18:41–50.

Tischendorf, L. and L. Fahrig, 2000a. How should we measure landscape connectivity?

Landscape Ecology 15:633–641.

Tischendorf, L. and L. Fahrig, 2000b. On the usage and measurement of landscape

connectivity. Oikos 90:7–19.

Treml, E. A., P. N. Halpin, D. L. Urban, and L. F. Pratson, 2008. Modeling population

connectivity by ocean currents, a graph-theoretic approach for marine conservation.

Landscape Ecology 23:19–36.

Turner, M. G., G. J. Arthaud, R. T. Engstrom, S. J. Hejl, J. Liu, S. Loeb, and

K. McKelvey, 1995. Usefulness of spatially explicit population models in land man-

agement. Ecological Applications 5:12–16. BIBLIOGRAPHY 171

Urban, D. and T. Keitt, 2001. Landscape connectivity: A graph-theoretic perspective.

Ecology 82:1205–1218.

Urban, D. L., 2005. Modeling ecological processes across scales. Ecology 86:19962006.

Urban, D. L., E. S. Minor, E. A. Treml, and R. S. Schick, 2009. Graph models of habi-

tat mosaics. Ecology Letters 12:260–273.

Van de Koppel, J., R. D. Bardgett, J. Bengtsson, C. Rodriguez-Barrueco, M. Rietkerk,

M. J. Wassen, and V. Wolters, 2005. The effects of spatial scale on trophic interac-

tions. Ecosystems 8:801–807.

Van der Heide, C. M., J. C. M. Van den Bergh, and E. C. Van Ierland, 2005. Extend-

ing Weitzman’s economic ranking of biodiversity protection: combining ecological

and genetic considerations. Ecological Economics 55:218–223.

Van Dyck, H. and M. Baguette, 2005. Dispersal behaviour in fragmented landscapes:

routine or special movements? Basic and Applied Ecology 6:535–545. van Langevelde, F., W. G. M. van der Knaap, and G. D. H. Claassen, 1998. Compar-

ing connectivity in landscape networks. Environment and Planning B: Planning and

Design 25:849–863.

Van Wagner, C. E., 1978. Age-class distribution and the forest fire cycle. Canadian

Journal of Forest Research 8:220–227.

Vasas, V., T. Magura, F. Jord`an,and B. Tothmeresz, 2009. Graph theory in action:

evaluating planned highway tracks based on connectivity measures. Landscape Ecol-

ogy 24:581–586.

Verbeylen, G., L. De Bruyn, F. Adriaensen, and E. Matthysen, 2003. Does matrix re-

sistance influence Red squirrel (Sciurus vulgaris L. 1758) distribution in an urban

landscape? Landscape Ecology 18:791–805. BIBLIOGRAPHY 172

Walker, B., C. S. Holling, S. R. Carpenter, and A. Kinzig, 2004. Resilience, adaptabil-

ity and transformability in socialecological systems. Ecology and Society 9:5.

Walker, W. and F. L. Craighead, 1997. Analyzing wildlife movement corridors in Mon-

tana using GIS. In Proceedings of the 1997 ESRI User Conference, pages 1–18.

Wasserman, S. and K. Faust, 1994. Social network analysis: methods and applications.

Cambridge University Press, New York, USA.

Watts, D. J., 1999. Small Worlds: The Dynamics of Networks between Order and Ran-

domness. Princeton University Press, Princeton.

Watts, D. J. and S. H. Strogatz, 1998. Collective dynamics of smallworld networks.

Nature 393:440–442.

Wiens, J., 1997. Metapopulation dynamics and landscape ecology. In I. Hanski and

M. E. Gilpin, editors, Metapopulation Biology, Ecology, Genetics, Evolution, pages

43–62. Academic Press, New York.

Wiens, J. A., 1989. Spatial scaling in ecology. Functional Ecology 3:385–397.

Wiens, J. A., 2001. The landscape context of dispersal. In J. Clobert, E. Danchin,

A. A. Dhondt, and J. D. Nichols, editors, Dispersal, pages 96–109. Oxford University

Press, Oxford.

Wiens, J. A., 2008. Allerton Park 1983: the beginnings of a paradigm for landscape

ecology? Landscape Ecology 23:125–128.

Williams, J. C., C. S. Re Velle, and S. Levin, 2004. Using mathematical optimization

models to design nature reserves. Frontiers in Ecology and the Environment 2:98–

105. BIBLIOGRAPHY 173

Williams, P., L. Hannah, S. Andelman, G. Midgley, M. Araujo, G. Hughes, L. Manne,

E. Martinez-Meyer, and R. Pearson, 2005. Planning for climate change: Identify-

ing minimum-dispersal corridors for the Cape proteaceae. Conservation Biology

19:1063–1074.

Williams, P. H. and M. B. Araujo, 2000. Using probability of persistence to identify

important areas for biodiversity conservation. Proceedings of the Royal Society of

London Series B-Biological Sciences 267:1959–1966.

Wimberly, M. C., 2006. Species dynamics in disturbed landscapes: when does a shift-

ing habitat mosaic enhance connectivity? Landscape Ecology 21:35–46.

Wintle, B. A., S. A. Bekessy, L. A. Veneir, and J. L. Pearce, 2005. The utility of dy-

namic landscape meta-population models for sustainable forest management: the

Brown Creeper. Conservation Biology 19:1930–1943.

With, K. A., 1994. Using fractal analysis to assess how species perceive landscape

structure. Landscape Ecology 9:25–36.

With, K. A., S. J. Cadaret, and C. Davis, 1999. Movement responses to patch struc-

ture in experimental fractal landscapes. Ecology 80:1340–1353.

With, K. A., R. H. Gardner, and M. G. Turner, 1997. Landscape connectivity and

population distributions in heterogeneous environments. Oikos 78:151–169.

With, K. A. and A. W. King, 1999. Dispersal success on fractal landscapes. A conse-

quence of lacuniarity thresholds. Landscape Ecology 14:73–82.

Witting, L., J. Tomiuk, and V. Loeschcke, 2000. Modelling the optimal conservation of

interacting species. Ecological Modelling 125:123–143.

Wright, S., 1943. Isolation by distance. Genetics 28:114–138. BIBLIOGRAPHY 174

Wu, J. and D. Marceau, 2002. Modeling complex ecological systems: an introduction.

Ecological Modelling 153:1–6.

Xu, W. H., Z. Ouyang, A. Vina, H. Zheng, J. G. Liu, and Y. Xiao, 2006. Designing a

conservation plan for protecting the habitat for giant pandas in the Qionglai moun-

tain range, China. Diversity and Distributions 12:610–619.

Yapp, R. H., 1922. The concept of habitat. The Journal of Ecology 10:1–17.