Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-rkxrd Total loading time: 0 Render date: 2024-07-18T16:21:34.806Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  31 October 2017

Mark R.T. Dale
Affiliation:
University of Northern British Columbia
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2017

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abarca-Arenas, L.G. & Ulanowicz, R.E. (2002). The effects of taxonomic aggregation on network analysis. Ecological Modelling, 149, 285296.Google Scholar
Abdi, R.H. (2007). RV coefficient and congruence coefficient. In Salkind, N. (ed.), Encyclopedia of Measurement and Statistics, pp. 849853. Thousand Oaks, CA: Sage.Google Scholar
Abrams, P.A., Menge, B.A., Mittelbach, G.G., Spiller, D.A. & Yodzis, P. (1996). The role of indirect effects in food webs. In Polis, G. & Winemuller, K.O. (eds), Food Webs: Integration of Patterns and Dynamics, pp. 371395. New York: Chapman & Hall.Google Scholar
Agresti, S., De Meo, P., Ferrara, E., Piccolo, S. & Provetti, A. (2015). Trust networks: Topology, dynamics, and measurements. IEEE Internet Computing, 19, 2635.Google Scholar
Ahmed, N.K., Neville, J. & Kompella, R. (2014). Network sampling: From static to streaming graphs. ACM Transactions on Knowledge Discovery from Data, 8, 156.Google Scholar
Ahmed, N., Neville, J., Rossi, R.A. & Duffield, N. (2015). Efficient graphlet counting for large networks. Proceedings of the IEEE International Conference on Data Mining, 110.Google Scholar
Ahn, Y.-Y., Bagrow, J.P. & Lehman, S. (2010). Link communities reveal multiscale complexity in networks. Nature, 466, 761764.Google Scholar
Ai, D., Gravel, D., Chu, C. & Wang, G. (2013). Spatial structures of the environment and of dispersal impact species distribution in competitive metacommunities. PLoS ONE, 8, e68927.Google Scholar
Airola, A., Pyysalo, S., Björne, J., Pahikkala, T., Ginter, F. & Salakoski, T. (2008). All-paths graph kernel for protein-protein interaction extraction with evaluation of cross-corpus learning. BioMed Central Bioinformatics, 9 (Suppl. 11): S2; doi:10.1186/1471-2105-9-S11-S2.Google Scholar
Albert, E.M., Fortuna, M.A., Godoy, J.A. & Bascompte, J. (2013). Assessing the robustness of networks of spatial genetic variations. Ecology Letters, 16, 8693.Google Scholar
Albert, R. & Barabási, A.-L. (2002). Statistical mechanics of complex networks. Reviews of Modern Physics, 74, 4797.Google Scholar
Alcántara, J.M. & Rey, P.J. (2012). Linking topological structure and dynamics in ecological networks. The American Naturalist, 180, 186199.Google Scholar
Alcántara, J.M. & Rey, P.J. (2014). Community dynamics: lessons from a skeleton. In Benítez, M., Miramontes, O. & Valiente-Banuet, A. (eds), Frontiers in Ecology, Evolution and Complexity, pp. 18 Mexico City: CopIt-arXives.Google Scholar
Alcántara, J.M., Rey, P.J. & Manzaneda, A.J. (2015). A model of plant community dynamics based on replacement networks. Journal of Vegetation Science, 26, 524537.Google Scholar
Alenazi, M.J.F. & Sterbenz, J.P.G. (2015). Comprehensive comparison and accuracy of graph metrics in predicting network resilience. Paper presented at the International Conference of Reliable Communication Networks, Kansas City, MO.Google Scholar
Alexandris, J. & Karagiorgos, G. (2014). Enhanced random walk with choice: an empirical study. International Journal on Applications of Graph Theory in Wireless ad hoc Networks and Sensor Networks, 6, doi:10.5121/jgraphoc.2014.6101.Google Scholar
Allesina, S., Bodini, A. & Bondavalli, C. (2005). Ecological subsystems via graph theory: the role of strongly connected components. Oikos, 110, 164176.Google Scholar
Allesina, S. & Levine, J.M. (2011). A competitive network theory of species diversity. Proceedings of the National Academy of Science, 108, 56385642. doi:10.1073/pnas.101442818.Google Scholar
Almeida-Neto, M., Guimaräes, P.R. & Lewinsohn, T.M. (2007). On nestedness analysis: rethinking matrix temperature and anti-nestedness. Oikos, 116, 716722.Google Scholar
Almeida-Neto, M., Guimaräes, P.R., Loyola, R.D. & Ulrich, W. (2008). A consistent metric for nestedness analysis in ecological systems: reconciling concept and measurement. Oikos, 117, 12271239.Google Scholar
Alon, N., Avin, C., Koucky, M., Kozma, G., Lotker, Z. & Tuttle, M.R. (2011). Many random walks are faster than one. Combinatorics, Probability and Computing, 20, 481502.Google Scholar
Altermatt, F., Seymour, M. & Martinez, N. (2013). River network properties shape α-diversity and community similarity patterns of aquatic insect communities across major drainage basins. Journal of Biogeography, 40, 22492260.Google Scholar
Amarasekare, P. (2009). Competition and coexistence in animal communities. In Levin, S.A. (ed.), The Princeton Guide to Ecology, pp. 196201. Princeton, NJ: Princeton University Press.Google Scholar
Anderson, T.K. & Sukhdeo, M.V.K. (2011). Host centrality in food web networks determines parasite diversity. PLoS ONE, 6 (10), e26798.Google Scholar
Andersson, E. & Bodin, Ö. (2009). Practical tool for landscape planning? An empirical investigation of network based models of habitat fragmentation. Ecography, 32, 123132.Google Scholar
Angelini, R., Aloísio, G.R. & Carvalho, A.R. (2010). Mixed food web control and stability in a Cerrado river (Brazil). Pan-American Journal of Aquatic Sciences, 10, 421431.Google Scholar
Arditi, R. & Michalski, J. (1996). Nonlinear food web models and their responses to increased basal productivity. In Polis, G. & Winemuller, K.O. (eds), Food Webs: Integration of Patterns and Dynamics, pp. 122133. New York: Chapman & Hall.Google Scholar
Atmar, W. & Patterson, B.D. (1993). The measure of order and disorder in the distribution of species in fragmented habitat. Oecologia, 96, 373382.Google Scholar
Avin, C., Koucký, M. & Lotker, Z. (2008). How to explore a fast-changing world (Cover time of a simple random walk on evolving graphs). In Aceto, L., Damgard, I., Goldberg, L.A., Halldórsson, M.M., Ingólfsdóttir, A. & Walukiewicz, I. (eds), Automata, Languages and Programming, pp. 121132 Berlin: Springer.Google Scholar
Avin, C. & Krishnamachari, B. (2008). The power of choice in random walks: an empirical study. Computer Networks, 52, 4460.Google Scholar
Azeria, E.T. (2004). Community Dynamics of Insular Biotas in Space and Time. PhD thesis, Swedish University of Agricultural Sciences, Uppsala.Google Scholar
Banašek-Richter, C., et al. (2009). Complexity in quantitative food webs. Ecology, 90, 14701477.Google Scholar
Baranyi, G., Saura, S., Podani, J. & Jordán, F. (2011). Contribution of habitat patches to network connectivity: redundancy and uniqueness of topological indices. Ecological Indicators, 11, 13011310.Google Scholar
Barbosa, P., Hines, J., Kaplan, I., Martinson, H., Szczepaniec, A. & Szendrei, Z. (2009). Associational resistance and associational susceptibility: having right or wrong neighbours. Annual Review of Ecology, Evolution and Systematics, 40, 120.Google Scholar
Barbour, A.D., Karónski, M. & Rucínski, A. (1989). A central limit theorem for decomposable random variables with applications to random graphs. Journal of Combinatorial Theory, Series B, 47, 125145.Google Scholar
Barcelo, H. & Laubenbacher, R. (2006). Graph-theoretic tools for dynamic network analysis. http://math.la.asu.edu/~helene/papers/blsocialjan06.pdf.Google Scholar
Bang-Jensen, J. & Gutin, G. (2009). Digraphs: Theory, Algorithms and Applications. 2nd ed. London: Springer.Google Scholar
Bascompte, J. (2007). Networks in ecology. Basic and Applied Ecology, 8, 485490.Google Scholar
Bascompte, J. (2009). Disentangling the web of life. Science, 325, 416419.Google Scholar
Bascompte, J. & Jordano, P. (2007). Plant-animal mutualistics networks: the architecture of biodiversity. Annual Review of Ecology, Evolution and Systematics, 38, 567593.Google Scholar
Bascompte, J. & Jordano, P. (2014). Mutualistic Networks. Princeton, NJ: Princeton University Press.Google Scholar
Bascompte, J., Jordano, P., Melián, C.J. & Olesen, J.M. (2003). The nested assembly of plant-animal mutualistic networks. Proceedings of the National Academy of Sciences, 100, 93Sep-8338.Google Scholar
Bascompte, J. & Stouffer, D.B. (2009). The assembly and disassembly of ecological networks. Philosophical Transactions of the Royal Society, Series B, 364, 17811787.Google Scholar
Baskerville, E.B., Dobson, A.P., Bedford, T., Allesina, S., Anderson, T.M. & Pascual, M. (2011). Spatial guilds in the Serengeti food web revealed by a Bayesian group model. PLoS Computational Biology, 7, e1002321; doi:10.1371/journalpcbi.1002321.Google Scholar
Bassett, D.S., Wymbs, N.F., Porter, M.A., Mucha, P.J., Carlson, J.M. & Grafton, S.T. (2011). Dynamic reconfiguration of human brain networks during learning. PNAS, 108, 76417646 doi:10.1073/pnas0.1018985108.Google Scholar
Bastolla, U., Fortuna, M.A., Pascual-Garcia, A., Ferrera, A., Luque, B. & Bascompte, J. (2009). The architecture of mutualistic networks minimizes competition and increases biodiversity. Nature, 458, 10181021; doi:10.1038/nature07950.Google Scholar
Beiler, K.J., Durall, D.M., Simard, S.W., Maxwell, S.A. & Kretzer, A.M. (2009). Architecture of the wood-wide web: Rhizopogon spp. genets link multiple Douglas-fir cohorts. New Phytologist, 185, 543553.Google Scholar
Belgrano, A., Scharler, U.M., Dunne, J. & Ulanowicz, R.E. (2005). Aquatic Food Webs: An Ecosystem Approach. Oxford: Oxford University Press.Google Scholar
Bell, A.D. (1974). Rhizome organization in relation to vegetative spread in Medeola virginiana. Journal of the Arnold Arboretum, 55, 458468.Google Scholar
Bellay, S., Lima, D.P., Takemoto, R.M. & Luque, J.L. (2011). A host-endoparasite network of neotropical marine fish: are there organizational patterns? Parasitology, 138, 19451952.Google Scholar
Benedek, Z., Jordán, F. & Báldi, A. (2007). Topological keystone species complexes in ecological interaction networks. Community Ecology, 8, 17.Google Scholar
Bengtsson, J., Angelstam, P., Elmqvist, T., Emanuelsson, U., Folke, C., Ihse, M., Moberg, F. & Nyström, M. (2003). Reserves, resilience, and dynamic landscapes. AMBIO, 32, 389396.Google Scholar
Benson, D.M., Grand, L.F., Vernia, C.S. & Gottwald, T.R. (2006). Temporal and spatial epidemiology of Phytophthora root rot in Fraser fir plantations. Plant Disease, 90, 11711180.Google Scholar
Beth, T., Borcherding, M. & Klein, B. (1994). Valuation of trust in open networks. Paper presented at the 3rd European Symposium on Research in Computer Security, Brighton, UK.Google Scholar
Bienert, A., Queck, R., Schmidt, A., Bernhofer, Ch. & Maas, H.-G. (2010). Voxel space analysis of terrestrial laser scans in forests for wind field modeling. International Archive for Photogrammetry, Remote Sensing, and Spatial Information Science, 38, 9297.Google Scholar
Biggs, N.L., Lloyd, E.K. & Wilson, R.J. (1976). Graph Theory 1736–1936. Oxford: Oxford University Press.Google Scholar
Bille, P. (2005). A survey on tree edit distance and related problems. Theoretical Computer Science, 337, 2234; doi:10.1016/j.tcs.2004.12.030.Google Scholar
Bjørnstadt, O.N., Stenseth, N.C. & Saitoh, T. (1999). Synchrony and scaling in dynamics of voles and mice in northern Japan. Ecology, 80, 622637.Google Scholar
Blanchet, F.G., Legendre, P. & Borcard, D. (2008). Modelling directional spatial processes in ecological data. Ecological Modelling, 215, 325336.Google Scholar
Blanchet, F.G., Legendre, P., Maranger, R., Monti, D. & Pepin, P. (2011). Modelling the effect of directional spatial ecological processes at different scales. Oecologia, 166, 357368.Google Scholar
Blick, R. & Burns, K.C. (2009). Network properties of arboreal plants: are epiphytes, mistletoes and lianas structured similarly? Perspectives in Plant Ecology, Evolution and Systematics, 11, 4152; doi:10.1016/j.ppees.2008.10.002.Google Scholar
Blonder, B. & Dornhaus, A. (2011). Time-ordered networks reveal limitations to information flow in ant colonies. PLoS ONE, 6 (5), e20298.Google Scholar
Blonder, B., Wey, T.W., Dornhaus, A., James, R. & Sih, A. (2012). Temporal dynamics and network analysis. Methods in Ecology and Evolution, 3, 958972; doi:10.1111/j.2041-210X.2012.00236.x.Google Scholar
Blüthgen, N., Fründ, J., Vazquez, D.P. & Menzel, F. (2008). What do interaction network metrics tell us about specialization and biological traits? Ecology, 89, 33873399.Google Scholar
Blüthgen, N., Menzel, F. & Blüthgen, N. (2006). Measuring specialization in species interaction networks. BMC Ecology, 6. doi:0.86/472-6785-6-9.Google Scholar
Bode, M., Burrage, K. & Possingham, H.P. (2008). Using complex network metrics to predict the persistence of metapopulations with asymmetric connectivity patterns. Ecological Modelling, 214, 201209.Google Scholar
Bonacich, P.F. (1972). Power and centrality: a family of measures. American Journal of Sociology, 92, 11701182.Google Scholar
Borcard, D., Gillet, F. & Legendre, P. (2011). Numerical ecology with R. New York: Springer Science.Google Scholar
Borcard, D., Legendre, P., Avois-Jacquet, C. & Tuomisto, H. (2004). Dissecting the spatial structure of ecological data at multiple scales. Ecology, 85, 18261832.Google Scholar
Borgatti, S.P. (2005). Centrality and network flow. Social Networks, 27, 5571.Google Scholar
Borgwardt, K.M. & Kriegel, H.-P. (2005). Shortest-path kernels on graphs. Paper presented at the International Conference on Data Mining, Houston, TX.Google Scholar
Borgwardt, K.M., Ong, C.S., Schönauer, S., Vishwanathan, S.V.N., Smola, A.J. & Kriegel, H.-P. (2005). Protein function prediction via graph kernels. Bioinformatics, 21 (s1), 110; doi:10.1093/bioinformatics/bti1007.Google Scholar
Borgwardt, K.M., Petri, T., Vishwanathan, S.V.N. & Kriegel, H.-P. (2007). An efficient sampling scheme for comparison of large graphs. Paper presented at the Conference on Mining and Learning with Graphs, Firence, Italy.Google Scholar
Boyland, N.K., James, R., Mlynski, D.T., Madden, J.R. & Croft, D.P. (2013). Spatial proximity loggers for recording animal social networks: consequences of inter-logger variation in performance. Behavioural and Ecological Sociobiology, 67, 18771890.Google Scholar
Brandes, U. & Fleischer, D. (2005). Centrality measures based on current flow. In Diekert, V. & Durand, B. (eds), Proceedings of the 22nd Annual Symposium on Theoretical Aspects of Computer Science, pp. 533544. Berlin: Springer.Google Scholar
Brin, S. & Page, L. (1998). The anatomy of a large-scale hypertextual Web search engine. Computer Networks & ISDN Systems, 30, 107117.Google Scholar
Brooks, C.P., Antonovics, J. & Keitt, T.H. (2008). Spatial and temporal heterogeneity explain disease dynamics in a spatially explicit network model. American Naturalist, 172, 1415-Sep9.Google Scholar
Brüggemann, R. & Carlsen, L. (2006). Partial Order in Environmental Sciences and Chemistry. Berlin: Springer.Google Scholar
Bryden, J., Funk, S., Geard, N., Bullock, S. & Jansen, V.A.A. (2011). Stability in flux: community structure in dynamic networks. Journal of the Royal Society Interface, 8, 10311040.Google Scholar
Buhl, J., Hicks, K., Miller, E.R., Persey, S., Alinvi, O. & Sumpter, D.J.T. (2009). Shape and effi-ciency of wood ant foraging networks. Behavioural Ecology & Sociobiology, 63, 451460.Google Scholar
Bukovinszky, T., van Veen, F.J.F., Jongema, Y. & Dicke, M. (2008). Direct and indirect effects of resource quality on food web structure. Science, 319, 804807.Google Scholar
Bunescu, R. & Mooney, R. (2005). A shortest path dependency kernel for relation extraction. In Proceedings of Human Language Technology Conference and Conference on Empirical Methods in Natural Language Processing, pp. 724731. Vancouver, Canada: Association for Computational Linguistics.Google Scholar
Bunn, A.G., Urban, D.L. & Keitt, T.H. (2000). Landscape connectivity: a conservation application of graph theory. Journal of Environmental Management, 59, 265278.Google Scholar
Burkle, L. & Irwin, R. (2009). The importance of interannual variation and bottom-up nitrogen enrichment for plant-pollinator networks. Oikos, 118. doi:10.1111/j.1600-0706.2009.17740.x.Google Scholar
Buss, L.W. & Jackson, J.B.C. (1979). Competitive networks: nontransitive competitive relationships in cryptic coral reef environments. American Naturalist, 113, 223234.Google Scholar
Butts, C.T. (2009). Revisiting the foundations of network analysis. Science, 325, 414416.Google Scholar
Caceres, R.S. & Berger-Wolf, T. (2013). Temporal scale of dynamic networks. In Holme, P. & Saramäki, J. (eds), Temporal Networks, pp. 6594. Berlin: Springer.Google Scholar
Cahill, J.F., Elle, E., Smith, G.R. & Shore, B.H. (2008). Disruption of a belowground mutualism alters interactions between plants and their floral visitors. Ecology, 89, 17911801.Google Scholar
Cahill, J.F., Kembel, S.W. & Gustafson, D.J. (2005). Differential genetic differences on competitive effect and response in Arabidopsis thaliana. Journal of Ecology, 93, 958967.Google Scholar
Canard, E.F., Mouquet, N., Mouillot, D., Stank, M., Miklisova, D. & Gravel, D. (2014). Empirical evaluation of neutral interactions in host-parasite networks. The American Naturalist, 183, 468479.Google Scholar
Cantor, M., Shoemaker, L.G., Cabral, R.B., Flores, C.O., Varga, M. & Whitehead, H. (2015). Multilevel animal societies can emerge from cultural transmission. Nature Communications, 6. doi:10.1038/ncomms9091.Google Scholar
Capitán, J.A., Arenas, A. & Guimerà, R. (2013). Degree of intervality of food webs: from body-size to models. Journal of Theoretical Biology, 334, 3544.Google Scholar
Cardinal, J., Collete, S. & Langerman, S. (2009). Empty region graphs. Computational Geometry: Theory and Applications, 42, 183195.Google Scholar
Careddu, G., Costantini, M.L., Calizza, E., Carlino, P., Bentivoglio, F., Orlandi, L. & Rossi, L. (2015). Effects of terrestrial input on macrobenthic food webs of a coastal sea are detected by stable isotope analysis in Gaeta Gulf. Estuarine, Coastal and Shelf Science, 154, 158168.Google Scholar
Carley, K.M. (2003). Dynamic network analysis. In Breiger, R., Carley, K.M. & Pattison, P. (eds), Dynamic Social Network Modeling and Analysis: Workshop Summary and Papers, pp. 133145. Washington, DC: Committee on Human Factors, National Research Council.Google Scholar
Carrara, F., Altermatt, F., Rodriguez-Iturbe, I. & Rinaldo, A. (2012). Dendritic connectivity controls biodiversity patterns in experimental metacommunities. Proceedings of the National Academy of Sciences, 15, 57615766.Google Scholar
Carrara, F., Rinaldo, A., Giometto, A. & Altermatt, F. (2014). Complex interaction of dendritic connectivity and hierarchical patch size on biodiversity in river-like landscapes. The American Naturalist, 183, 1325.Google Scholar
Carstensen, D.W., Sabatino, M., Trøjelsgaard, K. & Morellato, L.P.C. (2014). Beta diversity of plant-pollinator networks and the spatial turnover of pairwise interactions. PLoS ONE, 9 (11), e112903.Google Scholar
Cartozo, C.C., Garlaschelli, D. & Caldarelli, G. (2006). Graph theory and food webs. In Pascual, M. & Dunne, J.A. (eds), Ecological Networks: Linking Structure to Dynamics in Food Webs, pp. 93117. Oxford: Oxford University Press.Google Scholar
Cartwright, D. & Harary, F. (1956). Structural balance: a generalization of Heider's theory. Psychology Review, 63, 277293.Google Scholar
Casteigts, A., Flocchini, P., Quattrociocchi, W. & Santoro, N. (2011). Time-varying graphs and dynamic networks. arXiv:1012.0009v3 [cs.DC].Google Scholar
Cayley, A. (1857). On the theory of analytical forms called trees. Philosophical Magazine, 13, 1930.Google Scholar
Cerdeira, J.O., Pinto, L.S., Cabeza, M. & Gaston, K.J. (2010). Species specific connectivity in reserve-network design using graphs. Biological Conservation, 143, 408415.Google Scholar
Céréghino, R., Giraudel, J.L. & Compin, A. (2001). Spatial analysis of stream invertebrates distribution in the Adour-Garonne drainage basin (France), using Kohonen self-organization maps. Ecological Modelling, 146, 167180.Google Scholar
Chan, J., Bailey, J. & Leckie, C. (2008). Discovering correlated spatio-temporal changes in evolving graphs. Knowledge & Information Systems, 16, 5396.Google Scholar
Charbonneau, D., Blonder, B. & Dornhaus, A. (2013). Social insects: a model system for network dynamics. In Holme, X.P. & Saramäki, X.J. (eds), Temporal Networks, pp. 217244. Berlin: Springer.Google Scholar
Chardon, J.P., Adriaensen, F. & Matthysen, E. (2003). Incorporating landscape elements into a connectivity measure: a case study for the Speckled Wood Butterfly (Pararge aegeria L.). Landscape Ecology, 18, 561573.Google Scholar
Chartrand, G. & Lesniak, L. (2005). Graphs and Digraphs. 4th ed. Boca Raton, FL: Chapman & Hall/CRC.Google Scholar
Chave, J. (2009). Competition, neutrality, and community organization. In Levin, S.A. (ed.), The Princeton Guide to Ecology, pp. 264273. Princeton, NJ: Princeton University Press.Google Scholar
Chen, H.-W., Liu, W.-C., Davis, A.J., Jordán, F., Hwang, M.-J. & Shao, K.-T. (2008). Network position of hosts in food webs and their parasite diversity. Oikos, 117, 18471855.Google Scholar
Chetkiewicz, C.-L., St. Clair, C. & Boyce, M. (2006). Corridors for conservation: integrating pattern and process. Annual Review of Ecology, Evolution and Systematics, 37, 317342.Google Scholar
Chung, F.R.K. (1997). Spectral Graph Theory. Providence, RI: American Mathematical Society.Google Scholar
Clay, K. (1990). The impact of parasitic and mutualistic fungi on competitive interactions among plants. In Grace, J.B. & Tilman, D. (eds), Perspectives on Plant Competition, pp. 391412. San Diego, CA: Academic Press.Google Scholar
Codling, E.A., Plank, M.J. & Benhamou, S. (2008). Random walk models in biology. Journal of the Royal Society Interface, 5, 813834.Google Scholar
Cohen, J.E. & Palka, Z.J. (1990). A stochastic theory of community food webs. V. Intervality and triangulation in the trophic-niche overlap graph. The American Naturalist, 135, 435463.Google Scholar
Cohn, R.D. (1999). Comparisons of multivariate relational structures in serially correlated data. Journal of Agricultural, Biological & Environmental Statistics, 4, 238257.Google Scholar
Colinvaux, P.A. (1979). Why Big Fierce Animals Are Rare: An Ecologist's Perspective. Princeton, NJ: Princeton University Press.Google Scholar
Compieta, P., Di Martino, S., Bertolotto, M., Ferrucci, F. & Kechadi, T. (2007). Exploratory spatio-temporal data mining and visualization. Journal of Visual Languages & Computing, 18, 255279; doi:10.1016/j.jvc.2007.02.006.Google Scholar
Cook, R.E. (1988). Growth in Medeola virginiana clones. I. Field observations. American Journal of Botany, 75, 725731.Google Scholar
Cormen, T.H., Leiserson, C.E., Rivest, R.L. & Stein, C. (2009). Introduction to Algorithms. 3rd ed. Cambridge, MA: MIT Press.Google Scholar
Corson, F. (2010). Fluctuations and redundancy in optimal transport networks. Physical Review Letters, 104, 048703.Google Scholar
Cote, D., Kehler, D.G., Bourne, C. & Wiersma, Y.F. (2009). A new measure of longitudinal connectivity for stream networks. Landscape Ecology, 24, 101113.Google Scholar
Craft, M.E. & Caillaud, D. (2011). Network models: an underutilized tool in wildlife epidemiology. Interdisciplinary Perspectives on Infectious Diseases, 2011. doi:10.1155/2011/676949.Google Scholar
Craft, M.E., Volz, E., Packer, C. & Meyers, L.A. (2010). Disease transmission in territorial populations: the small-world network of Serengeti lions. Journal of the Royal Society Interface, 8. doi:10.1098/rsif.2010.0511.Google Scholar
Cressie, N.A.C. (1993) Statistics for Spatial Data. 2nd ed. New York: John Wiley.Google Scholar
Cressie, N. & Wikle, C.K. (2011). Statistics for Spatio-Temporal Data. Hoboken, NJ: John Wiley.Google Scholar
Croft, D.P., Madden, J.R., Franks, D.W. & James, R. (2011). Hypothesis testing in animal social networks. Trends in Ecology and Evolution, 26, 502507.Google Scholar
Cui, S., Hero, A.O., Luo, Z.-Q. & Moura, J.M.F. (2016). Big Data over Networks. Cambridge: Cambridge University Press.Google Scholar
Cumming, G.S. (2016). Heterarchies: Reconciling networks and hierarchies. Trends in Ecology and Evolution, XX31. doi.org/10.1016/j.tree.2016.04.009.Google Scholar
Dale, M.R.T. (1977a). Graph theoretical analysis of the phytosociological structure of plant communities: the theoretical basis. Vegetatio, 34, 137154.Google Scholar
Dale, M.R.T. (1977b). Graph theoretical analysis of the phytosociological structure of plant communities: an application to mixed forest. Vegetatio, 35, 3546.Google Scholar
Dale, M.R.T. (1982). Phytosociological structure of seaweed communities and the invasion of Fucus serratus in Nova Scotia. Canadian Journal of Botany, 60, 26522658.Google Scholar
Dale, M.R.T. (1984). The contiguity of upslope and downslope boundaries of species in a zoned community. Oikos, 42, 9296.Google Scholar
Dale, M.R.T. (1986). Overlap and spacing of species’ ranges on an environmental gradient. Oikos, 47, 303308.Google Scholar
Dale, M.R.T. (1999). Spatial Pattern Analysis in Plant Ecology. Cambridge: Cambridge University Press.Google Scholar
Dale, M.R.T. & Fortin, M.-J. (2009). Spatial autocorrelation and statistical tests: some solutions. Journal of Agricultural, Biological, and Environmental Statistics, 14, 188206.Google Scholar
Dale, M.R.T. & Fortin, M.-J. (2010). From graphs to spatial graphs. Annual Review of Ecology, Evolution, and Systematics, 41, 2138.Google Scholar
Dale, M.R.T. & Fortin, M.-J. (2014). Spatial Analysis. 2nd ed. Cambridge: Cambridge University Press.Google Scholar
Dale, M.R.T., John, E.A. & Blundon, D.J. (1991). Contact sampling for the detection of interspecific association: a comparison in two vegetation types. Journal of Ecology, 79, 781792.Google Scholar
Dale, M.R.T. & Moon, J. (1988). Statistical tests on two characteristics of the shapes of cluster diagrams. Journal of Classification, 5, 2138.Google Scholar
Dale, M.R.T. & Thomas, A.G. (1987). The structure of weed communities in Saskatchewan fields. Weed Science, 35, 348355.Google Scholar
Danon, L., Ford, A.P., House, T., Jewell, C.P., Keeling, M.J., Roberts, G.O., Ross, J.V. & Vernon, M.C. (2011). Networks and the epidemiology of infectious disease. Interdisciplinary Perspectives on Infectious Diseases, 2011. doi:10.1155/2011/284909.Google Scholar
Darlington, P.J. (1957). Zoogeography: The Geographical Distribution of Animals. New York: John Wiley.Google Scholar
Darwin, C. (1862). On the Various Contrivances by Which British and Foreign Orchids Are Fertilized by Insects and on the Good Effects of Intercrossing. London: John Murray.Google Scholar
David, F.N., Kendall, M.G. & Barton, D.E. (1966). Symmetric Functions and Allied Tables. Cambridge: Cambridge University Press.Google Scholar
Davis, J.A. (1967). Clustering and structural balance in graphs. Human Relations, 29, 181187.Google Scholar
Davis, K.F., D'Odorico, P., Laio, F. & Ridolfi, L. (2013). Global spatio-temporal patterns in human migration: a complex network perspective. PLoS ONE, 8, e53723; doi:10.1371/journal.pone.0053723.Google Scholar
Decout, S., Manel, S., Miaud, C. & Luque, S. (2012). Integrative approach for landscape based graph connectivity analysis: a case study with the common frog (Rana temporaria), in human dominated landscapes. Landscape Ecology, 27, 267279.Google Scholar
de Jong, P., Aarssen, L.W. & Turkington, R. (1983). The analysis of contact sampling data. Journal of Ecology, 71, 545559.Google Scholar
Del Mondo, G., Stell, J.G., Claramunt, C. & Thibaud, R. (2010). A graph model for spatio-temporal evolution. Journal of Universal Computer Science, 16, 14521477.Google Scholar
Demattei, C. & Cucala, L. (2009). Multiple spatio-temporal cluster detection for case event data: an ordering-based approach. Communications in Statistics: Theory & Methods, 40, 3 Mar-5872.Google Scholar
den Boer, P.J. (1968). Spreading of risk and stabilization of animal numbers. Acta Biotheoretica, 18, 165194.Google Scholar
Deshpande, M., Kuramochi, M., Wale, N. & Karypis, G. (2005). Frequent substructure-based approaches for classifying chemical compounds. IEEE Transactions on Knowledge and Data Engineering, 17, 10361050.Google Scholar
de Vries, H., Stevens, J.M.G. & Vervaecke, H. (2006). Measuring and testing the steepness of dominance hierarchies. Animal Behaviour, 71, 585592.Google Scholar
Di Battista, G., Eades, P., Tamassia, R. & Tollis, I.G. (1994). Algorithms for drawing graphs. Computational Geometry, 4, 235282.Google Scholar
Ding, C., He, X., Xiong, H., Peng, H. & Holbrook, S.R. (2006). Transitive closure and metric inequality of weighted graphs: detecting protein interaction modules using cliques. International Journal of Data Mining and Informatics, 1, 162177.Google Scholar
Do, A.-L.J. & Gross, T. (2012). Self-Organization in Continuous Adaptive Networks. Aalborg, Denmark: River.Google Scholar
Donetti, L., Neri, F. & Munoz, M.A. (2006). Optimal network topologies: expanders, cages, Ramanujan graphs, entangled networks and all that. Journal of Statistical Mechanics, 8, P 8007; arXiv:cond-mat/0605565v2.Google Scholar
Doreian, P. & Mrvar, A. (1996). A partitioning approach to structural balance. Social Networks, 18, 149168.Google Scholar
Dormann, C.F. (2006). Competition hierarchy, transitivity and additivity: investigating the effect of fertilization on plant-plant interactions using three common bryophytes. Plant Ecology, 191, 171184.Google Scholar
Dormann, C.F. (2011). How to be a specialist? Quantifying specialization in pollination networks. Network Biology, 1, 120.Google Scholar
Dormann, C.F., Fründ, J., Blüthgen, N. & Gruber, B. (2009). Indices, graphs, and null models: analyzing bipartite ecological networks. The Open Ecological Journal, 2, 724.Google Scholar
Dorogovstsev, S.N. & Mendes, J.F.F. (2002). Evolution of networks. Advances in Physics, 51, 10791187.Google Scholar
Dos Santos, D.A., Fernandez, H.R., Cuezzo, M.G. & Dominguez, E. (2008). Sympatry inference and network analysis in biogeography. Systematic Biology, 57, 432448.Google Scholar
Doyle, F.I. & Smith, J.N.M. (2001). Raptors and scavengers. In Krebs, C.J., Boutin, S. & Boonstra, R. (eds), Ecosystem Dynamics of the Boreal Forest: The Kluane Project, pp. 377404 Oxford: Oxford University Press.Google Scholar
Dray, S., Chessel, D. & Thioulouse, J. (2003). Co-inertia analysis and the linking of the ecological data tables. Ecology, 84, 30783089.Google Scholar
Dray, S., Legendre, P. & Peres-Neto, P.R. (2006). Spatial modelling: a comprehensive framework for principal coordinate analysis of neighbour matrices (PCNM). Ecological Modelling, 196, 483493.Google Scholar
Dray, S., Pélissier, R., Couteron, P., Fortin, M.-J., Legendre, P., Peres-Neto, P.R., Bellier, E., et al. (2012). Community ecology in the age of multivariate multiscale spatial analysis. Ecological Monographs, 82, 257275.Google Scholar
Drewe, J.A., Weber, N., Carter, S.P., Bearhop, S., Harrison, X.A., Dall, S.R.X., McDonald, R.A. & Delahay, R.J. (2012). Performance of proximity loggers in recording intra- and inter-species interactions: a laboratory and field-based validation study. PLoS ONE, 7, e39068.Google Scholar
Duboscq, J., Romano, V., Sueur, C. & MacIntosh, A.J.J. (2015). Network centrality and seasonality interact to predict lice load in a social primate. Scientific Reports, 6, 22095; doi:10.1038/srep22095.Google Scholar
Duncan, A.J., Gunn, G.J., Lewis, F.I., Umstatter, C. & Humphry, R.W. (2012). The influence of empirical contact networks on modelling diseases in cattle. Epidemics, 4, 117123.Google Scholar
Dunne, J.A. (2006). The network structure of food webs. In Pascual, M. & Dunne, J.A. (eds), Ecological Networks: Linking Structure to Dynamics in Food Webs, pp. 2786. Oxford: Oxford University Press.Google Scholar
Dunne, J.A., Lafferty, K.D., Dobson, A.P., Hechinger, R.F. & Kuris, A.M. (2013). Parasites affect food web structure primarily through increased diversity and complexity. PLoSBiol, 11 (6): e1001579; doi:10.1371/journal.pbio.1001579.Google Scholar
Dunne, J.A., Williams, R.J. & Martinez, N.D. (2002a). Network structure and biodiversity loss in food-webs: robustness increases with connectance. Ecology Letters, 5, 558567.Google Scholar
Dunne, J.A., Williams, R.J. & Martinez, N.D. (2002b). Food-web structure and network theory: the role of connectance and size. Proceedings of the National Academy of Science, 99, 1291712922.Google Scholar
Durrett, R. (2007). Random Graph Dynamics. Cambridge: Cambridge University Press.Google Scholar
Dyer, R.J., Chan, D.M., Gardiakos, V.A. & Meadows, C.A. (2012). Pollination graphs: quantifying pollen pool covariance networks and the influence of intervening landscape on genetic connectivity in the North American understory tree, Cornus florida L. Landscape Ecology, 27, 239251.Google Scholar
Dyer, R.J. & Nason, J.D. (2004). Population graphs: the graph theoretic shape of genetic structure. Molecular Ecology, 13, 17131727.Google Scholar
Economo, E.P. & Keitt, T.H. (2008). Species diversity in neutral metacommunities: a network approach. Ecology Letters, 11, 5262.Google Scholar
Edgington, E.S. (1995). Randomization Tests. 3rd ed. New York: Marcel Dekker.Google Scholar
Efron, B. & Tibshirani, R.J. (1993). An Introduction to the Bootstrap. New York: Chapman & Hall.Google Scholar
Eifrem, E. (2014). Graph theory: key to understanding big data. Wired, May 5.Google Scholar
Eiler, A. (2006). Evidence for the ubiquity of mixotrophic bacteria in the upper ocean: implications and consequences. Applied and Environmental Microbiology, 72 (12), 74317437; doi:10.1128/AEM.01559-06.Google Scholar
Erdös, P. (1963). On a problem in graph theory. Mathematical Gazette, 47, 220223.Google Scholar
Erdös, P., Harary, F. & Tutte, W.T. (1965). On the dimension of a graph. Mathematika, 12, 118122.Google Scholar
Erdös, P. & Moon, J.W. (1964). On subgraphs of the complete bipartite graph. Canadian Mathematical Bulletin, 7, 3539.Google Scholar
Erdös, P. & Rényi, A. (1960). On the evolution of random graphs. Publications of the Mathematical Institute of the Hungarian Academy of Sciences, 5, 1761.Google Scholar
Erös, T., Olden, J.D., Schick, R.S., Schmera, D. & Fortin, M.-J. (2012). Characterizing connectivity relationships in freshwaters using patch-based graphs. Landscape Ecology, 27, 303317.Google Scholar
Escoufier, Y. (1973). Le traitement des variables vectorielles. Biometrics, 29, 751760.Google Scholar
Escudero, A., Pajarón, S. & Gavilán, R. (1994). Saxicolous communities in the Sierra del Moncayo (Spain): a classificatory study. Coenoses, 9, 1524.Google Scholar
Estrada, E. (2006). Network robustness to targeted attacks: the interplay of expansibility and robustness. The European Physical Journal B, 52, 563574.Google Scholar
Estrada, E. (2007). Characterization of topological keystone species: local, global and “meso-scale” centralities in food webs. Ecological Complexity, 4, 4657.Google Scholar
Estrada, E. (2012). The Structure of Complex Networks: Theory and Applications. Oxford: Oxford University Press.Google Scholar
Estrada, E. & Bodin, Ö. (2008). Using network centrality measures to manage landscape connectivity. Ecological Applications, 18, 18101825.Google Scholar
Facchetti, G., Iacono, G. & Altafini, C. (2011). Computing global structural balance in large-scale signed social networks. Proceedings of the National Academy of Sciences, 108, 2095320958.Google Scholar
Fagan, W.F. (2002). Connectivity, fragmentation, and extinction risk in dendritic metapopulations. Ecology, 83, 32433249.Google Scholar
Fagan, W.F. & Calabrese, J.M. (2014). The correlated random walk and the rise of movement ecology. Bulletin of the Ecological Society of America, 95, 204206.Google Scholar
Fall, A., Fortin, M.-J., Manseau, M. & O'Brien, D. (2007). Spatial graphs: principles and applications for habitat connectivity. Ecosystems, 10, 448461.Google Scholar
Fang, W. (2005). Spatial analysis of an invasion front of Acer platanoides: dynamic inferences from static data. Ecography, 28, 283294.Google Scholar
Farine, D.R. & Whitehead, H. (2015). Constructing, conducting and interpreting animal social network analysis. Journal of Animal Ecology, 84, 11441163.Google Scholar
Fattorini, S. (2007). Non-randomness in the species-area relationship: testing the underlying mechanisms. Oikos, 116, 678689.Google Scholar
Fattorini, S. & Fowles, A.P. (2005). A biogeographical analysis of the tenebrionid beetles (Coleoptera, Tenebrionidae) of the island of Thasos in the context of the Aegean Islands (Greece). Journal of Natural History, 39, 39193949.Google Scholar
Fedor, A. & Vasas, V. (2009). The robustness of keystone indices in food webs. Journal of Theoretical Biology, 260, 372378.Google Scholar
Feeley, K.J., Gillespie, T.W., Lebbin, D.J. & Walter, H.S. (2007). Species characteristics associated with extinction vulnerability and nestedness rankings of birds in tropical forest fragments. Animal Conservation, 10, 493501.Google Scholar
Ferrari, J.R., Lookingbill, T.R. & Neel, M.C. (2007). Two measures of landscape graph connectivity: assessments across gradients in area and configuration. Landscape Ecology, 22, 1315; doi:10.1007/s10980-007-9121-7.Google Scholar
Flitcroft, R.L., Burnett, K.M., Reeves, G.H. & Ganio, L.M. (2012). Do network relationships matter? Comparing network and instream habitat variables to explain densities of juvenile coho salmon (Oncorhynchus kisutch) in mid-coastal Oregon, USA. Aquatic Conservation: Marine and Freshwater Ecosystems, 22, 288302.Google Scholar
Fodor, E. (2013). Linking biodiversity to mutualistic networks: woody species and ectomycorrhizal fungi. Annals of Forest Research, 56, 5378.Google Scholar
Fortin, M.-J. & Payette, S. (2002). How to test the significance of the relation between spatially autocorrelated data at the landscape scale: a case study using fire and forest maps. Ecoscience, 9, 213218.Google Scholar
Fortuna, M.A. & Bascompte, J. (2008). The network approach in ecology. In Valladares, F., Camacho, A., Elosegui, A., Estrada, M. & Garcia, C. (eds), Unity in Diversity: Reflections on Ecology after the Legacy of Ramon Margalef, pp. 371–93. Madrid: Fundación BBVA.Google Scholar
Fortuna, M.A., Garcia, C., Guimaräes, P.R. & Bascompte, J. (2008). Spatial mating networks in insect-pollinated plants. Ecology Letters, 11, 490498.Google Scholar
Fortuna, M.A., Gómez-Rodriguez, C. & Bascompte, J. (2006). Spatial network structure and amphibian persistence in stochastic environments. Proceedings of the Royal Society B: Biological Sciences, 273, 14291434.Google Scholar
Fortuna, M.A., Popa-Lisseanu, A.G., Ibañaz, C. & Bascompte, J. (2009). The roosting spatial network of a bird-predator bat. Ecology, 90, 934944.Google Scholar
Fortunato, S. (2010). Community detection in graphs. arXiv:0906.0612v2 [physics.soc-ph].Google Scholar
Fowler, N.L. (1990). Disorderliness in plant communities: comparisons, causes, and consequences. In Grace, J.B. & Tilman, D. (eds), Perspectives on Plant Competition, pp. 291306. San Diego, CA: Academic Press.Google Scholar
Franc, A. (2003). Metapopulation dynamics as a contact process on a graph. Ecological Complexity, 1, 4963; doi:10.1016/jecocom.2003.10.02.Google Scholar
Frank, K. (1995). Identifying cohesive subgroups. Social Networks, 17, 2756.Google Scholar
Frean, M. & Abraham, E.R. (2001). Rock-scissors-paper and the survival of the weakest. Proceedings of the Royal Society of London B, 268, 13231327.Google Scholar
Freckleton, R.P. & Watkinson, A.R. (2001). Predicting competition coefficients for plant mixtures: reciprocity, transitivity and correlations with life-history traits. Ecology Letters, 4, 348357.Google Scholar
Freeman, L.C., Borgatti, S.P. & White, D.R. (1991). Centrality in valued graphs: a measure of betweenness based on network flow. Social Networks, 13, 141154.Google Scholar
Fuller, M.M., Wagner, A. & Enquist, B.J. (2008). Using network analysis to characterize forest structure. Natural Resource Modelling, 21, 225247.Google Scholar
Galpern, P., Manseau, M. & Fall, A. (2011). Patch-based graphs of landscape connectivity: a guide to construction, analysis and application for conservation. Biological Conservation, 144, 4455.Google Scholar
Ganguly, N., Deutsch, A. & Mukherjee, A. (2009). Dynamics on and of Complex Networks. Boston: Birkhäuser.Google Scholar
Ganio, L.M., Torgersen, C.E. & Gresswell, R.E. (2005). A geostatistical approach for describing spatial pattern in stream networks. Frontiers in Ecology and Environment, 3, 138144.Google Scholar
Gao, X., Xiao, B., Tao, D. & Li, X. (2010). A survey of graph edit distance. Pattern Analysis and Applications, 13, 113129; doi:10.1007/s10044-008-0141-y.Google Scholar
Garcia, D., Zamora, R. & Amico, G.C. (2011). The spatial scale of plant-animal interactions: effects of resource availability and habitat structure. Ecological Monographs, 81, 103121.Google Scholar
Garroway, C.J., Bowman, J., Carr, D. & Wilson, P.J. (2008). Applications of graph theory to landscape genetics. Evolutionary Applications, 1, 620630.Google Scholar
Gärtner, T. (2003). A survey of kernels for structured data. ACM SIGKDD Explorations Newsletter, 5 (1), 4958.Google Scholar
Gärtner, T. (2008). Kernels for Structured Data. Singapore: World Scientific.Google Scholar
Gärtner, T., Flach, P.A. & Wrobel, S. (2003). On graph kernels: hardness results and efficient alternatives. In Proceedings of the Sixteenth Annual Conference on Learning Theory and Seventh Annual Workshop on Kernel Machines, Lecture Notes in Artificial Intelligence, pp. 129143. New York: Springer.Google Scholar
Gasalla, M.A., Rodrigues, A.R. & Postuma, F.A. (2010). The trophic role of the squid Loligo plei as a keystone species in the South Brazil Bight ecosystem. ICES Journal of Marine Science, 67, 14131424.Google Scholar
Gehring, C.A., Mueller, R.C., Haskins, K.E., Rubow, T.K. & Whitham, T.G. (2014). Convergence in mycorrhizal fungal communities to drought, plant competition, parasitism, and susceptibility to herbivory: consequences for fungi and host plants. Frontiers in Microbiology, 5, 306; doi:10.3389/micb.2014.00306.Google Scholar
Gilarranz, L.J. & Bascompte, J. (2012). Spatial network structure and metapopulation persistence. Journal of Theoretical Biology, 297, 1116.Google Scholar
Gilarranz, L.J., Hastings, A. & Bascompte, J. (2015). Inferring topology from dynamics in spatial networks. Theoretical Ecology, 8 doi:10.1007/s12080-014-0231-y.Google Scholar
Girvan, M. & Newman, M.E.J. (2002). Community structure in social and biological networks. Proceedings of the National Academy of Sciences, 99, 78217826.Google Scholar
Glaz, J., Naus, J. & Wallenstein, S. (2001). Scan Statistics. New York: Springer.Google Scholar
Gleick, J. (1987). Chaos: The Making of a New Science. New York: Viking.Google Scholar
Godfrey, S.S. (2013). Networks and the ecology of parasite transmission: a framework for wildlife parasitology. International Journal for Parasitology: Parasites and Wildlife, 2, 235245.Google Scholar
Godfrey, S.S., Moore, J.A., Nelson, N.J. & Bull, C.M. (2010). Social network structure and parasite infection patterns in a territorial reptile, the Tuatara (Sphenodon punctatus). International Journal of Parasitology, 40, 15751585.Google Scholar
Gollo, L.L. & Breakspear, M. (2014). The frustrated brain: from dynamics on motifs to communities and networks. Philosophical Transactions of the Royal Society, Series B, 369; doi:10.1098/rstb.2013.0532.Google Scholar
Gómez, J.M., Perfectii, F. & Jordano, P. (2011). The functional consequences of mutualistic network architecture. PLoS ONE, 6, e16143; doi.10.1371/journal.pone.0016143.Google Scholar
Gómez, S., Montiel, J., Torres, D. & Fernández, A. (2012). MultiDendrograms: variable-group agglomerative hierarchical clusterings. arXiv:1201.1623v2.Google Scholar
Goodall, D.W. (1966). The nature of the mixed community. Proceedings of the Ecological Society of Australia, 1, 8496.Google Scholar
Gordon, A.D. & Birks, H.J.B. (1972). Numerical methods in Quaternary paleoecology. I. Zonation of pollen diagrams. New Phytologist, 71, 961979.Google Scholar
Gordon, A.D. & Birks, H.J.B. (1974). Numerical methods in Quaternary paleoecology. II. Comparison of pollen diagrams. New Phytologist, 73, 221249.Google Scholar
Grace, J.B., Guntespurgen, G.R. & Keough, J. (1993). The examination of a competition matrix for transitivity and intransitive loops. Oikos, 68, 9198.Google Scholar
Graham, S.P., Hassan, H.K., Burkett-Cadena, N.D., Guyer, C. & Unnasch, T.R. (2009). Nestedness of ectoparasite-vertebrate host networks. PLoS ONE, 4, e7873; doi:10.1371/journal.pone.0007873.Google Scholar
Grant, E.H.C., Lowe, W.H. & Fagan, W.F. (2007). Living in the branches: population dynamics and ecological processes in dendritic networks. Ecology Letters, 10, 165175.Google Scholar
Greenland, S., Pearl, J. & Robins, J.M. (1999). Causal diagrams for epidemiological research. Epidemiology, 10, 3748.Google Scholar
Grime, J.P. (1979). Plant Strategies and Vegetation Processes. New York: John Wiley.Google Scholar
Grindrod, P., Parsons, M.C., Higham, D.J. & Estrada, E. (2011). Communicability across evolving networks. Physical Review E, 83, 046120; doi:10.1103/PhysRevE.83.046120.Google Scholar
Gross, T. & Blasius, B. (2008). Adaptive coevolutionary networks: a review. Journal of the Royal Society Interface, 5, 259271.Google Scholar
Guerrero, C., Milenkovic, T., Pržulj, N., Kaiser, P. & Huang, L. (2008). Characterization of the proteasome interaction network using a qtax-based tag-team strategy and protein interaction network analysis. Proceedings of the National Academy of Sciences, 105, 1333313338.Google Scholar
Guha, R., Kumar, K., Raghavan, P. & Tomkins, A. (2004). Propagation of trust and distrust. Paper ACM 1-58113-844-X/04/0005 presented at ACM WWW, New York.Google Scholar
Guillaume, J.-L. & Latapy, M. (2006). Bipartite graphs as models of complex networks. Physica A, 371, 795813.Google Scholar
Guimerà, R. & Sales-Pardo, M. (2006). Form follows function: the architecture of complex networks. Molecular Systems Biology, 2 Article 42. doi:10.1038/msb4100082.Google Scholar
Guimerà, R. & Sales-Pardo, M. (2009). Missing and spurious interactions and the reconstruction of complex networks. Proceeding of the National Academy of Sciences, 52, 2207322078.Google Scholar
Hagen, M., et al. (2012). Biodiversity, species interactions and ecological networks in a fragmented world. Advances in Ecological Research, 46, 89185.Google Scholar
Hamede, R.K., Bashford, J., McCallum, H. & Jones, M. (2009). Contact networks in a wild Tasmanian devil (Sarcophilus harrisii) population: using social network analysis to reveal seasonal variability in social behavior and its implications for transmission of devil facial tumour disease. Ecology Letters, 12, 111.Google Scholar
Hammer, A.C. & Pitchford, J.W. (2005). The role of mixotrophy in plankton bloom dynamics and the consequences for productivity. ICES Journal of Marine Science, 62, 833840.Google Scholar
Hampton, S.E., Strasser, C.A., Tewksbury, J.J., Gram, W.K., Budden, A.E., Batcheller, A.L., Duke, C.S. & Porter, J.H. (2013). Big data and the future of ecology. Frontiers of Ecology and Environment, 11, 156162; doi:10.1890/120103.Google Scholar
Hanski, I. (2009). Metapopulations and spatial population processes. In Levin, S.A. (ed.), The Princeton Guide to Ecology, pp. 177185. Princeton, NJ: Princeton University Press.Google Scholar
Harary, F. (1953). On the notion of balance of a signed graph. Michigan Mathematical Journal, 2, 143146.Google Scholar
Harary, F. (1959). On the measurement of structural balance. Behavioral Science, 4, 316323.Google Scholar
Harary, F. (1969). Graph Theory. Reading MA: Addison-Wesley.Google Scholar
Harary, F. & Gupta, G (1997). Dynamic graph models. Mathematical & Computer Modelling, 25, 7987.Google Scholar
Harper, J.L. (1977). Population Biology of Plants. London: Academic Press.Google Scholar
Harvey, E. & Miller, T.E. (1996). Variance in composition of inquiline communities in leaves of Sarracenia purpurea L. on multiple spatial scales. Oecologia, 108, 562566.Google Scholar
Hastie, T., Tibshirani, R. & Friedman, J. (2009). The Elements of Statistical Learning: Data Mining, Inference, and Prediction. 2nd ed. New York: Springer.Google Scholar
Hobaiter, C., Poisot, T., Zuberbühler, K., Hoppitt, W. & Gruber, T. (2014). Social network analysis shows direct evidence for social transmission of tool use in wild chimpanzees. PLOS Biology, 12, e1001960.Google Scholar
Hock, K. & Mumby, P.J. (2015). Quantifying the reliability of dispersal paths in connectivity networks. Journal of the Royal Society Interface, 12, 20150013; doi:10.1098/rsif.2015.0013.Google Scholar
Hodges, K.E., Krebs, C.J., Hik, D.S., Stefan, C.I., Gillis, E.A. & Doyle, C.E. (2001). Snowshoe hare demography. In Krebs, C.J., Boutin, S. & Boonstra, R. (eds), Ecosystem Dynamics of the Boreal Forest: The Kluane Project, pp. 141178 Oxford: Oxford University Press.Google Scholar
Hofmann, T., Schölkopf, B. & Smola, A.J. (2008). Kernel methods in machine learning. The Annals of Statistics, 36, 11711220; doi:10.1214/009053607000000677.Google Scholar
Hollén, L.I. & Radford, A.N. (2009). The development of alarm call behavior in mammals and birds. Animal Behaviour, 78, 791800.Google Scholar
Holme, P. (2013). Epidemiologically optimal static networks from temporal network data. PLoS Computational Biology, 9, e1003142.Google Scholar
Holme, P. & Saramäki, J. (2011). Temporal networks. Physics Reports, 519, 97125.Google Scholar
Holme, P. & Saramäki, J. (2013). Temporal Networks. Berlin: Springer.Google Scholar
Holt, R.D. (2002). Food webs in space: on the interplay of dynamic in stability and spatial processes. Ecological Research, 17, 261273.Google Scholar
Holt, R.D. (2009). Predation and community organization. In Levin, S.A. (ed.), The Princeton Guide to Ecology, pp. 274281. Princeton, NJ: Princeton University Press.Google Scholar
Holt, R.D. & Hoopes, M.F. (2005). Food web dynamics in a metacommunity context: modules and beyond. In Holyoak, P., Leibold, M.A. & Holt, R.D. (eds), Metacommunities: Spatial Dynamics and Ecological Communities, pp. 6893. Chicago: Chicago University Press.Google Scholar
Holyoak, M. (2000). Habitat subdivision causes changes in food web structure. Ecology Letters, 3, 509515.Google Scholar
Horn, H.S. (1975). Markovian processes of forest succession. In Cody, M.L. & Diamond, J.M. (eds), Ecology and Evolution of Communities, pp. 196211. Cambridge, MA: Belknap Press.Google Scholar
Horn, H.S. (1976). Succession. In May, R.M. (ed.), Theoretical Ecology: Principles and Applications, pp. 187204. Philadelphia: Saunders.Google Scholar
Horváth, D.X. & Kertész, J. (2014). Spreading dynamics on networks: the role of burstiness, topology and non-stationarity. New Journal of Physics, 16. doi:10.1088/1367-2630/16/7/73037.Google Scholar
Howard, T.G. (2001). The relationship of total and per-gram rankings in competitive effect to the natural abundance of herbaceous perennials. Journal of Ecology, 89, 110117.Google Scholar
Huffaker, C.B. (1958). Experimental studies on predation: dispersion factors and predator-prey oscillations. Hilgardia, 27, 343383.Google Scholar
Hulovatyy, Y., Chen, H. & Milenkovic, T. (2015). Exploring the structure and function of temporal networks with dynamic graphlets. Bioinformatics, 31, 11711180.Google Scholar
Husté, A. & Boulinier, T. (2011). Determinants of bird community composition on patches in suburbs of Paris, France. Biological Conservation, 144, 243252.Google Scholar
Ilany, A., Barocas, A., Koren, L., Kam, M. & Geffen, E. (2013). Structural balance in the social networks of a wild mammal. Animal Behaviour, 85, 13971405.Google Scholar
Illian, J.B., Penttinen, A., Stoyan, H. & Stoyan, D. (2008). Statistical Analysis and Modelling of Spatial Point Patterns. New York: John Wiley.Google Scholar
Ings, T.C., et al. (2008). Ecological networks – beyond food webs. Journal of Animal Ecology, 78. doi:10.1111/j.1365-2656.2008.01460.x.Google Scholar
Itzhaki, Z., Akiva, E. & Margalit, H. (2010). Preferential use of protein domain pairs as interaction mediators: order and transitivity. Bioinformatics, 26, 25642570.Google Scholar
Ives, A.R. & Godfray, H.C.J. (2006). Phylogenetic analysis of trophic associations. The American Naturalist, 168, E1E14; www.jstor.org/stable/10.1086/505157.Google Scholar
Iyengar, V.S. (2005). Space-time clusters with flexible shapes. Morbidity and Mortality Weekly Report, 54, 7176.Google Scholar
Jacquez, G.M. (1996). A k nearest neighbour test for space–time interaction. Statistical Medicine, 15, 19351949.Google Scholar
James, A., Pitchford, J.W. & Plank, M.J. (2012). Disentangling nestedness from models of ecological complexity. Nature, 487, 227230.Google Scholar
James, P., Rayfield, B., Fortin, M.-J., Fall, A. & Farley, G. (2005). Reserve network design combining spatial graph theory and species’ spatial requirements. Geomatica, 59, 323333.Google Scholar
James, P.M.A., Fleming, R.A. & Fortin, M.-J. (2010). Identifying significant scale-specific spatial boundaries using wavelets and null models: spruce budworm defoliation in Ontario, Canada, as a case study. Landscape Ecology, 25, 873887.Google Scholar
James, P.M.A., Fortin, M.-J., Fall, A., Kneeshaw, D. & Messier, C. (2007). The effects of spatial legacies following shifting management practices and fire on boreal forest age structure. Ecosystems, 10, 12611277.Google Scholar
James, R., Croft, D.P. & Krause, J. (2009). Potential banana skins in animal social network analysis. Behavioural Ecology and Sociobiology, 63, 989997.Google Scholar
Jelínková, H., Tremblay, F. & DesRochers, A. (2009). Molecular and dendrochronological analyses of natural root grafting in Populus tremuloides (Salicaceae). American Journal of Botany, 96, 15001505; doi:10.3732/ajb.0800177.Google Scholar
Jiang, L.Q. & Zhang, W.J. (2015). Determination of keystone species in CSM food web: A topological analysis of network structure. Network Biology, 5, 1333.Google Scholar
Johnson, S., Domínguez-García, V. & Muñoz, M.A. (2013). Factors determining nestedness in complex networks. PLoS ONE, 8 doi:10.1371/journal.pone.0074025.Google Scholar
Jones, D.A. (1966). On the polymorphism of cyanogenesis in Lotus corniculatus: Selection by animals. Canadian Journal of Genetics and Cytology, 8, 556567.Google Scholar
Joppa, L.N., Bascompte, J., Montoya, J.M., Solé, R.V., Sanderson, J. & Pimm, S.L. (2009). Reciprocal specialization in ecological networks. Ecology Letters, 12, 961969; doi:101111/j.1461-0248.2009.01341.x.Google Scholar
Jordán, F. (2009). Keystone species and food webs. Philosophical Transactions of the Royal Society, Series B, 364, 17331741.Google Scholar
Jordán, F., Liu, W.-C. & Davis, A.J. (2006). Topological keystone species: measures of positional importance in food webs. Oikos, 112, 535546.Google Scholar
Jordán, F., Takács-Sánta, A. & Molnár, I. (1999). A reliability theoretical quest for keystones. Oikos, 86, 453462.Google Scholar
Jordano, P. (1987). Patterns of mutualistic interactions in pollination and seed dispersal: connectance, dependence asymmetries, and coevolution. American Naturalist, 129, 657677.Google Scholar
Jordano, P. (2010). Coevolution in multispecific interactions among free-living species. Evolution, Education, and Outreach, 3, 4046.Google Scholar
Jordano, P., Bascompte, J. & Olesen, J.M. (2003). Invariant properties in coevolutionary networks of plant-animal interactions. Ecology Letters, 6, 6981.Google Scholar
Jordano, P., Bascompte, J. & Olesen, J.M. (2006). The ecological consequences of complex topology and nested structure in pollination webs. In Waser, N.M. & Olleton, J. (eds), Plant-Pollinator Interactions, pp. 173199. Chicago: Chicago University Press.Google Scholar
Juenger, M. & Mutzel, P (2003). Graph Drawing Software. Berlin: Springer.Google Scholar
Kampichler, C., Angeler, D.G., Holmes, R.T., Leito, A., Svensson, S., van der Jeungd, H.P. & Wesolowski, T. (2014). Temporal dynamics of bird community composition: an analysis of baseline conditions from long-term data. Oecologia, 175, 13011313; doi:10.1007/s00442-014-2979-6.Google Scholar
Karsai, M., Kivelä, M., Pan, R.K., Kaski, K., Kertész, J., Barabási, A.L. & Saramäki, J. (2011). Small but slow world: how network topology and burstiness slow down spreading. Physical Review E, 83, 025102. arXiv:1006.2125v3.Google Scholar
Katz, L. (1953). A new status index derived from sociometric analysis. Psychometrica, 18, 3943.Google Scholar
Keats, D.W., South, G.R. & Steele, D.H. (1985). Algal biomass and diversity in the upper subtidal at a pack-ice disturbed site in eastern Newfoundland. Marine Ecology Progress Series, 25, 151163.Google Scholar
Keddy, P.A. & Shipley, B. (1989). Competitive hierarchies in herbaceous plant communities. Oikos, 54, 234241.Google Scholar
Keeling, M.J. (1999). The effects of local spatial structure on epidemiological invasions. Proceedings of the Royal Society of London, Series B, 266, 859867.Google Scholar
Keeling, M.J. (2005). The implications of network structure for epidemic dynamics. Theoretical Population Biology, 67, 18.Google Scholar
Keeling, M.J. & Eames, K.T.D. (2005). Networks and epidemic models. Journal of the Royal Society Interface, 2, 295307; doi:10.1098/rsif.2005.0051.Google Scholar
Kéfi, S., Berlow, E.L., Wieters, E.A., Joppa, L.N., Wood, S.A., Brose, U. & Navarrete, S.A. (2015). Network structure beyond food webs: mapping non-trophic and trophic interactions on Chilean rocky shores. Ecology, 96, 291303.Google Scholar
Keitt, T.H., Urban, D.L. & Milne, B.T. (1997). Detecting critical scales in fragmented landscapes. Conservation Ecology, 1, 4. www.consecol.org/vol1/iss1/art4/.Google Scholar
Kempe, D., Kleinberg, J. & Kumar, A. (2002). Connectivity and inference problems for temporal networks. Journal of Computer Systems Science, 64, 820842.Google Scholar
Kent, M. & Coker, P. (1992). Vegetation Description and Analysis. Boca Raton, FL: CRC Press.Google Scholar
Kepner, J., Ricke, D.O. & Hutchison, D. (2013). Taming biological big data with D4M. Lincoln Laboratory Journal, 20, 8291.Google Scholar
Kernighan, B.W. & Lin, S. (1970). An efficient heuristic procedure for partitioning graphs. Bell System Technical Journal, 49, 291307.Google Scholar
Kim, H. & Anderson, R. (2012). Temporal node centrality in complex networks. Physical Review E, 85, 026107.Google Scholar
Kirchhoff, G. (1847). Über die Auflösung der Gleichungen. Annals of Physical Chemistry, 72, 497508.Google Scholar
Kirkpatrick, D.G. & Radke, J.D. (1985). A framework for computational morphology. Computational Geometry, Machine Intelligence and Pattern Recognition, 2, 217248.Google Scholar
Kisilevich, S., Mansmann, F., Nanni, M. & Rinzivillo, S. (2010). Spatio-temporal clustering: a survey. In Maimon, O. & Rokach, L. (eds), Data Mining and Knowledge Discovery Handbook, pp. 855874. New York: Springer.Google Scholar
Kitching, R.L. (2000). Food Webs and Container Habitats. Cambridge: Cambridge University Press.Google Scholar
Klovdahl, A.S. (1985). Social networks and the spread of infectious diseases: the AIDS example. Social Science and Medicine, 21, 12031218.Google Scholar
Kneitel, J.M. & Miller, T.E. (2003). Dispersal rates affect species composition in metacommunities of Sarracenia purpurea L. inquilines. American Naturalist, 162, 165171.Google Scholar
Kocarev, L. & Vattay, G. (2010). Complex Dynamics in Communication Networks. Berlin: Springer.Google Scholar
Kolaczyk, E.D. (2009). Statistical Analysis of Network Data. New York: Springer.Google Scholar
Kolaczyk, E.D. & Csárdi, G. (2014). Statistical Analysis of Network Data with R. New York: Springer.Google Scholar
Kondoh, M., Kato, S. & Sakato, Y. (2010). Food webs are built up with nested subwebs. Ecology, 91, 31233130.Google Scholar
Kondor, R.I. & Lafferty, J. (2002). Diffusion kernels on graphs and other discrete input spaces. Proceedings of the International Conference on Machine Learning (ICML).Google Scholar
Kondor, R., Shervashidze, N. & Borgwardt, K.M. (2009). The graphlet spectrum. Paper presented at the 26th International Conference on Machine Learning, Montreal.Google Scholar
Kost, C., de Oliveira, E.G., Knoch, T.A. & Wirth, R. (2005). Spatio-temporal permanence and plasticity of foraging trails in young and mature leaf-cutting ant colonies (Atta sp.). Journal of Tropical Ecology, 21, 677688.Google Scholar
Kostakos, V. (2009). Temporal graphs. Physica A, 388, 10071023.Google Scholar
Kovanen, L., Karsai, M., Kaski, K., Kertész, J. & Saramäki, J. (2011). Temporal motifs in time-dependent networks. Journal of Statistical Mechanics, 11, P11005; arXiv:1107.5646v2.Google Scholar
Kovanen, L., Karsai, M., Kaski, K., Kertész, J. & Saramäki, J. (2013a). Temporal motifs. In Holme, P. & Saramäki, J. (eds), Temporal Networks, pp. 119134. Berlin: Springer.Google Scholar
Kovanen, L., Kaski, K., Kertesz, J. & Saramäki, J. (2013b). Temporal motifs reveal homophily, gender-specific patterns, and group talk in call sequences. Proceedings of the National Academy of Sciences, 110, 1807018075.Google Scholar
Krause, A.E., Frank, K.A., Mason, D.M., Ulanowicz, R.E. & Taylor, W.W. (2003). Compartments revealed in food-web structure. Nature, 426, 282285.Google Scholar
Krause, J., Croft, D.P. & James, R. (2007). Social network theory in the behavioural sciences. Behavioural Ecology and Sociobiology, 62, 1527.Google Scholar
Krebs, C.J. (2010). Of lemmings and snowshoe hares: the ecology of northern Canada. Proceedings of the Royal Society, Series B, 278 doi:10.1098/rspb.2010.1992.Google Scholar
Krebs, C.J., Boutin, S. & Boonstra, R. (2001). Ecosystem dynamics of the boreal forest: the Kluane project. Oxford: Oxford University Press.Google Scholar
Krishna, A., Guimaräes, P.R., Jordano, P. & Bascompte, J. (2008). A neutral-niche theory of nestedness in mutualistic networks. Oikos, 117, 16091618; doi:10.1111/j.2008.0030-1299.16540.x.Google Scholar
Kunegis, J. (2014). Applications of structural balance in signed social networks. arXiv:1402.6865v1 [cs.S1].Google Scholar
Kunegis, J., Schmidt, S., Lommatzsch, A., Lerner, J., De Luca, E.W. & Albayrak, S. (2010). Spectral analysis of signed graphs for clustering, prediction, and visualization. In Proceedings of the 10th SIAM International Conference on Data Mining, pp. 559570. SIAM: Columbus, Ohio.Google Scholar
Kurvers, R.H.J.M., Krause, J., Croft, D.P., Wilson, A.D.M. & Wolf, M. (2014). The evolutionary and ecological consequences of animal social networks: emerging issues. Trends in Ecology and Evolution, 29, 326335Google Scholar
Labonne, J., Ravigné, V., Parisi, B. & Gaucherel, C. (2008). Linking dendritic network structures to population demogenetics: the downside of connectivity. Oikos, 117, 14791490.Google Scholar
Laird, R.A. & Schamp, B.S. (2006). Competitive intransitivity promotes species coexistence. American Naturalist, 168, 182193.Google Scholar
Laird, R.A. & Schamp, B.S. (2008). Does local competition increase the coexistence of species in intransitive networks? Ecology, 89, 237247.Google Scholar
Laita, A., Kotiaho, J.S. & Mönkkönen, M. (2011). Graph-theoretic connectivity measures: what do they tell us about connectivity? Landscape Ecology, 26, 951967.Google Scholar
Lamour, A., Termorshuizen, A.J., Volker, D. & Jeger, M.J. (2007). Network formation by rhizomorphs of Armillaria lutea in natural soil: their description and ecological significance. FEMS Microbial Ecology, 62, 222232.Google Scholar
Landau, H.G. (1951). On dominance relations and the structure of animal societies: I. Effects of inherent characteristics. Bulletin of Mathematica Biophysics, 13, 119.Google Scholar
Lankau, R.A., Wheeler, E., Bennett, A.E. & Strauss, S.Y. (2010). Plant-soil feedbacks contribute to an intransitive competitive network that promotes genetic and species diversity. Journal of Ecology, 99 doi 10.1111/j.1365-2745.2010.01736.x.Google Scholar
Lapointe, F.-J. & Legendre, P. (1990). A statistical framework to test the consensus of two nested classifications. Systematic Zoology, 39, 113.Google Scholar
Lapointe, F.-J. & Legendre, P. (1995). Comparison tests for dendrograms: a comparative evaluation. Journal of Classification, 12, 265282.Google Scholar
Lawson, A.B. (2013). Bayesian Disease Mapping: Hierarchical Modeling in Spatial Epidemiology. Boca Raton, FL: CRC Press.Google Scholar
Layeghifard, M., Makarenkov, V. & Peres-Neto, P.R. (2015). Spatial and species compositional networks for inferring connectivity patterns in ecological communities. Global Ecology and Biogeography, 24, 718727.Google Scholar
Leal, W., Llanos, E.J., Retrepo, G., Suarez, C.F. & Patarrayo, M.E. (2016). How frequently do clusters occur in hierarchical clustering analysis? A graph theoretical approach to studying ties in proximity. Journal of Cheminformatics, 8 doi:10.1186/s13321-016-0114-x.Google Scholar
Lee, S.H., Kim, P.-J. & Jeong, H. (2006). Statistical properties of sampled networks. Physical Review E, 73 016102.Google Scholar
Legendre, P. (1993). Spatial autocorrelation: trouble or new paradigm? Ecology, 74, 16591673.Google Scholar
Legendre, P. & Fortin, M.-J. (1989). Spatial pattern and ecological analysis. Vegetatio, 80, 107138.Google Scholar
Legendre, P. & Legendre, L. (2012). Numerical Ecology. 3rd ed. Amsterdam: Elsevier.Google Scholar
Leibold, M.A. (2009). Spatial and metacommunity dynamics in biodiversity. In Levin, S.A. (ed.), The Princeton Guide to Ecology, pp. 312319. Princeton, NJ: Princeton University Press.Google Scholar
Leskovec, J. & Faloutsos, C. (2006). Sampling from large graphs. In Proceedings of the 12th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, pp. 631636. ACM: New York, USAGoogle Scholar
Leskovec, J., Huttenlocher, D. & Kleinberg, J. (2010). Signed networks in social media. Paper ACM 978-1-60558-929-9/10/04 presented at ACM CHI, Atlanta, GA, USA.Google Scholar
Lesne, A. (2006). Complex networks: from graph theory to biology. Letters in Mathematical Physics, 78, 235262.Google Scholar
Levine, J.M. (1999). Indirect facilitation: evidence and predictions from a riparian community. Ecology, 80, 17621769.Google Scholar
Levinton, J.S. (1982). Marine Ecology. Engelwood Cliffs, NJ: Prentice Hall.Google Scholar
Libralato, S., Christensen, V. & Pauly, D. (2006). A method for identifying keystone species in food web models. Ecological Modelling, 195, 153171.Google Scholar
Liebhold, A., Koenig, W.D. & Bjørnstadt, O.N. (2004). Spatial synchrony in population dynamics. Annual Review of Ecology, Evolution & Systematics, 35, 467490.Google Scholar
Lihoreau, M., Raine, N.E., Reynolds, A.M., Stelzer, R.J., Lim, K.S., Smith, A.D., Osborne, J.L. & Chittka, L. (2013). Unravelling the mechanisms of trapline foraging in bees. Communicative & Integrative Biology, 6, e22701.Google Scholar
Little, L.R. & Dale, M.R.T. (1999). A method for analyzing spatio-temporal pattern in plant establishment, tested on a Populus balsamifera clone. Journal of Ecology, 87, 620627.Google Scholar
Liu, W., Li, D., Wang, J., Xie, H., Zhu, Y. & He, F. (2009). Proteome-wide prediction of signal flow direction in protein interaction networks based on interacting domains. Molecular and Cellular Proteomics, 8, 20632070.Google Scholar
Loewe, K., Grueschow, M., Stoppel, C.M., Kruse, R. & Borgelt, C. (2014). Fast construction of voxel-level functional connectivity graphs. BMC Neuroscience, 15, 78; doi:10.1186/1471-2202-15-78.Google Scholar
Lomolino, M.V. (2006). Investigating causality of nestedness in insular communities: selective immigrations or extinctions. Journal of Biogeography, 23, 699703.Google Scholar
Loreau, M. (2010). From Populations to Ecosystems. Princeton, NJ: Princeton University Press.Google Scholar
Lorenz, K. (1991). Here Am I – Where Are You? New York: Harcourt.Google Scholar
Lovász, L. (1993). Random walks on graphs: a survey. In Combinatorics: Paul Erdös Is Eighty, pp. 146 Bolyai Mathematical Studies, vol. 2. Budapest: Keszthely.Google Scholar
Lugo-Martinez, J. & Radivojac, P. (2014). Generalized graphlet kernels for probabilistic inference in sparse graphs. Network Science, 2, 254276.Google Scholar
Lundberg, J. & Moberg, F. (2003). Mobile link organisms and ecosystem functioning: implications for ecosystem resilience and management. Ecosystems, 6, 8798.Google Scholar
Lundgren, J.R. & Maybee, J.S. (1984). Food webs with interval competition graphs. In Graphs and Applications: Proceedings of the First Colorado Symposium on Graph Theory, pp. 231244. New York: John Wiley.Google Scholar
Luque, S., Saura, S. & Fortin, M.-J. (2012). Landscape connectivity analysis for conservation: insights from combining new methods with ecological and genetic data. Landscape Ecology, 27, 153157.Google Scholar
Maiya, A.S. (2011). Sampling and Inference in Complex Networks. PhD thesis, Department of Computer Science, University of Illinois at Chicago.Google Scholar
Malliaros, F.D. & Vazirgiannis, M. (2013). Clustering and community detection in directed networks: a survey. Physics Reports, 533, 95142; arXiv:1308.0971v1 [cs.SI].Google Scholar
Manly, B.F.J. (2006). Randomization, Bootstrap, and Monte Carlo Methods in Biology. 3rd ed. Boca Raton, FL: Chapman & Hall/CRC.Google Scholar
Mantzaris, A.V. & Higham, D.J. (2013). Dynamic communicability predicts infectiousness. In Holme, P. & Saramäki, J. (eds), Temporal Networks, pp. 283294. Berlin: Springer.Google Scholar
Marsh, M.K., McLeod, S.R., Hutchings, M.R. & White, P.C.L. (2011). Use of proximity loggers and network analysis to quantify social interactions in free-ranging wild rabbit populations. Wildlife Research, 38, 112.Google Scholar
Marsten, R. (2014). Is graph theory key to understanding big data? Wired, March.Google Scholar
Martinez, N.D. (1991). Artifacts or attributes? Effects of resolution on the Little Rock Lake food web. Ecological Monographs, 61, 367392.Google Scholar
Martinez, N.D. (1992). Constant connectance in community foodwebs. American Naturalist, 139, 12081218.Google Scholar
Mason, O. & Verwoerd, M. (2007). Graph theory and networks in biology. IET Systems Biology, 1, 89119.Google Scholar
May, R.M. (2006). Network structure and the biology of populations. Trends in Ecology and Evolution, 21. doi:10.1016/tree.2006.03.013.Google Scholar
Maynard-Smith, J. (1982). Evolution and the Theory of Games. Cambridge: Cambridge University Press.Google Scholar
McCann, K. (2009). The structure and stability of food webs. In Levin, S.A. (ed.), The Princeton Guide to Ecology, pp. 305311. Princeton, NJ: Princeton University Press.Google Scholar
McCann, K. (2012). Food Webs. Princeton, NJ: Princeton University Press.Google Scholar
McIntire, E.J.B. & Fajardo, A. (2009). Beyond description: the active and effective way to infer processes from spatial patterns. Ecology, 90, 4656.Google Scholar
McQuaid, C.F. & Britton, N.F. (2013). Network dynamics contribute to structure: nestedness in mutualistic networks. Bulletin of Mathematical Biology, 75, 23722388.Google Scholar
McRae, B.H. & Beier, P. (2007). Circuit theory predicts gene flow in plant and animal populations. Proceedings of the National Academy of Sciences, 10, 1988519890.Google Scholar
McRae, B.H., Dickson, B.G., Keitt, T.H. & Shah, V.B. (2008). Using circuit theory to model connectivity in ecology, evolution, and conservation. Ecology, 89, 27122724.Google Scholar
Medan, D., Perazzo, R.P.J., Devoto, M., Burgos, E., Zimmermann, M.G., Ceva, H. & Delbue, A.M. (2007). Analysis and assembling of network structure in mutualistic systems. Journal of Theoretical Biology, 246, 510521.Google Scholar
Melián, C.J. & Bascompte, J. (2004). Food web cohesion. Ecology, 85, 352358.Google Scholar
Melián, C.J., Bascompte, J., Jordano, P. & Křivan, V. (2009). Diversity in a complex ecological network with two interaction types. Oikos, 118, 122130.Google Scholar
Menge, B.A. (1995). Indirect effects in a marine rocky intertidal interaction webs: patterns and importance. Ecological Monographs, 65, 2174.Google Scholar
Meyers, L.A. (2006). Predicting epidemics on directed contact networks. Journal of Theoretical Biology, 240, 400418.Google Scholar
Meyers, L.A. (2007). Contact network epidemiology: bond percolation applied to infectious disease prediction and control. Bulletin of the American Mathematical Society, 44, 6386.Google Scholar
Miele, V., Picard, F. & Dray, S. (2014). Spatially constrained clustering of ecological networks. Methods in Ecology and Evolution, 5, 771779.Google Scholar
Mihail, J.D. & Bruhn, J.N. (2005). Foraging behavior of Armillaria rhizomorph systems. Mycological Research, 109, 11951207.Google Scholar
Milenkovic, T. (2008). Graph-Theoretical Approaches for Studying Biological Networks. PhD thesis, Department of Computer Science, University of California, Irvine.Google Scholar
Milenkovic, T. & Pržulj, N. (2008). Uncovering biological network function via graphlet degree signatures. Cancer Informatics, 6, 257273.Google Scholar
Milo, R., Shen-Orr, S., Itzkovitz, S., Kashtan, N., Chklovskii, D. & Alon, U. (2002). Network motifs: simple building blocks of complex networks. Science, 298, 824827.Google Scholar
Min, B. & Goh, K.I. (2013). Burstiness: measures, models, and dynamic consequences. In Holme, P. & Saramäki, J. (eds), Temporal Networks, pp. 4164. Berlin: Springer.Google Scholar
Minor, E.S. & Urban, D.L. (2007). Graph theory as a proxy for spatially explicit population models in conservation planning. Ecological Applications, 17, 17711782.Google Scholar
Minor, E.S. & Urban, D.L. (2008). A graph-theory framework for evaluating landscape connectivity and conservation planning. Conservation Biology, 22, 297307.Google Scholar
Mitchell, R.J. (1992). Testing evolutionary and ecological hypotheses using path analysis and structural equation modelling. Functional Ecology, 6, 123129.Google Scholar
Mladenoff, D. (1999). Spatial Modeling of Forest Landscape Change. Cambridge: Cambridge University Press.Google Scholar
Mnih, V., et al. (2015). Human-level control through deep reinforcement learning. Nature, 518, 529533; doi:10.1038/nature14236.Google Scholar
Mnih, V., Kavukcuoglu, K., Silver, D., Graves, A., Antonoglu, I., Wierstra, D. & Riedmiller, M. (2013). Playing Atari with deep reinforcement learning. ArXiv 1312.5602.Google Scholar
Moilanen, A. (2011). On the limitations of graph-theoretic connectivity in spatial ecology and conservation. Journal of Applied Ecology, 48, 15431547; doi:10.1111/j.1365-2664.2011.02062.x.Google Scholar
Montoya, J.M., Pimm, S.L. & Solé, R.V. (2006). Ecological networks and their fragility. Nature, 442, 259264.Google Scholar
Moody, J. (2008). Static Representations of Dynamic Networks. Durham, NC: Duke Population Research Institute.Google Scholar
Moon, J.W. (1968). Topics on Tournaments. New York: Holt, Rinehart, Winston.Google Scholar
Morin, P.J. & Lawler, S.P. (1996). Effects of food chain length and omnivory on population dynamics in experimental food webs. In Polis, G. & Winemuller, K.O. (eds), Food Webs: Integration of Patterns and Dynamics, pp. 122133. New York: Chapman & Hall.Google Scholar
Morris, R.J., Gripenberg, S., Lewis, O.T. & Roslin, T. (2014). Antagonistic interaction networks are structured independently of latitude and host guild. Ecology Letters, 17, 340349.Google Scholar
Mortelliti, A., Westgate, M., Stein, J., Wood, J., & Lindenmayer, D.B. (2015). Ecological and spatial drivers of population synchrony in bird assemblages. Basic & Applied Ecology, 16, 269278.Google Scholar
Moschitti, A. (2006). Making tree kernels practical for natural language processing. Paper presented at the Conference of the European Chapter of the Association for Computational Linguistics, Trento, Italy.Google Scholar
Mouillot, D., Krasnov, B.R. & Poulin, R. (2008). High intervality explained by phylogenetic constraints in host-parasite webs. Ecology, 89, 20432051.Google Scholar
Mouquet, N. & Loreau, M. (2002). Coexistence in metacommunities: the regional similarity hypothesis. American Naturalist, 159, 420426.Google Scholar
Mouritsen, K.N., Poulin, R., McLaughlin, J.P. & Thieltges, D.W. (2011). Food web including metazoan parasites for an intertidal ecosystem in New Zealand. Ecology Archives, 92, 92173; doi:10.1890/11-0371.1.Google Scholar
Mucha, H.-J., Bartel, H.-G. & Dolata, J. (2005). Techniques of rearrangements in binary trees (dendrograms) and applications. MATCH (Communications in Mathematical and Computer Chemistry), 54, 561582.Google Scholar
Mucha, P.J., Richardson, T., Macon, K., Porter, M.A. & Onnela, J.-P. (2010). Community structure in time-dependent, multiscale, and multiplex networks. Science, 328, 876878; doi:10.1126/science.1184819.Google Scholar
Murphy, M.A., Dezzani, R.D.S., Pilliod, D.S. & Storfer, A. (2010). Landscape genetics of high mountain frog metapopulations. Molecular Ecology, 19, 36343649.Google Scholar
Murtagh, F. (1984). Counting dendrograms: a survey. Discrete Applied Mathematics, 17, 191199.Google Scholar
Nacher, J.C., Ueda, N., Yamada, T., Kanehisa, M. & Akutsu, T. (2004). Clustering under the line graph transformation: application to reaction network. BioMed Central Bioinformatics, 5, 207; doi:10.1186/1471-2105-5-207.Google Scholar
Naujokaitis-Lewis, I.R., Rico, Y., Lovell, J., Fortin, M.-J. & Murphy, M.A. (2013). Implications of incomplete networks on estimation of landscape genetic connectivity. Conservation Genetics, 14, 287298.Google Scholar
Neutel, A.M., Heesterneek, J.A.P. & de Ruiter, P.C. (2002). Stability in real food webs: weak links in long loops. Science, 296, 11201123.Google Scholar
Newman, M.E.J. (2002). Assortative mixing in networks. Physics Review Letters, 89, 208701.Google Scholar
Newman, M.E.J. (2006). Modularity and community structure in networks. Proceedings of the National Academy of Sciences, 103, 85778582.Google Scholar
Newman, M.E.J. (2010). Networks: An Introduction. Oxford: Oxford University Press.Google Scholar
Newman, M.E.J., Barabási, A.-L. & Watts, D., eds (2006). The Structure and Dynamics of Networks. Princeton, NJ: Princeton University Press.Google Scholar
Newman, M.E.J. & Girvan, M. (2004). Finding and evaluating community structure in networks. Physics Review E, 69, 026113; arXiv: cond-mat/0308217v1.Google Scholar
Newman, M.E.J. & Park, J. (2003). Why social networks are different from other types of networks. Physics Review E, 68, 036122; arXiv:cond-mat/0305612v1.Google Scholar
Newman, M.E., Watts, D.J. & Strogatz, S.H. (2002). Random graph models of social networks. PNAS, 99, 25662572.Google Scholar
Nguyen-Phuc, B.T. & Dale, M.R.T. (2014). Sub-graphs in competitive hierarchies. Unpublished discussion manuscript, University of Alberta.Google Scholar
Nicosia, V., Tang, J., Mascolo, C., Musolesi, M. & Latora, V. (2013). Graph metrics for temporal networks. In Holme, P. & Saramäki, J. (eds), Temporal Networks, pp. 1540. Berlin: Springer.Google Scholar
Nicosia, V., Tang, J., Musolesi, M., Mascolo, C., Russo, G. & Latora, V. (2012). Components in time-varying graphs. Chaos, 22, 023101; arXiv: 1106.2134v3.Google Scholar
O'Brien, D., Manseau, M., Fall, A. & Fortin, M.-J. (2006). Testing the importance of spatial configuration of winter habitat for woodland caribou: an application of graph theory. Biological Conservation, 130, 7083.Google Scholar
Oh, S.J., Joung, J.-G., Chang, J.-H. & Zhang, B.-T. (2006). Construction of phylogenetic trees by kernel-based comparative analysis of metabolic networks. BioMed Central Bioinformatics. doi:10.1186/1471-2105-7-284.Google Scholar
Ohgushi, T. (2008). Herbivore-induced indirect interaction webs in terrestrial plants: the importance of non-trophic, indirect and facilitative interactions. Entomologia Experimentalis et Applicata, 128, 217229.Google Scholar
Okabe, A., Boots, B., Sugihara, K. & Chiu, S.N. (2000). Spatial Tessellations: Concepts and Applications of Voronoi Diagrams. 2nd ed. Chichester: Wiley.Google Scholar
Okabe, A. & Sugihara, K. (2012). Spatial Analysis along Networks. New York: John Wiley.Google Scholar
Okey, T.A. & Pauly, D. (1999). Trophic mass-balance model of Alaska's Prince William Sound ecosystem, for the post-spill period 1994–1996. University of British Columbia, Vancouver.Google Scholar
Olesen, J.M., Bascompte, J., Dupont, Y.L., Elberling, H., Rasmussen, C. & Jordano, P. (2011b). Missing and forbidden links in mutualistic networks. Proceedings of the Royal Society, Series B, 278, 725732.Google Scholar
Olesen, J.M., Bascompte, J., Elberling, H. & Jordano, P. (2008). Temporal dynamics in a pollination network. Ecology, 89, 15731582.Google Scholar
Olesen, J.M., Stefanescu, C. & Traveset, A. (2011a). Strong long-term temporal dynamics of an ecological network. PLoS ONE, 11, e26455.Google Scholar
Openshaw, S. (1984). The Modifiable Area Unit Problem. Norwich: Geo Books.Google Scholar
Osborne, J.L., Clark, S.J., Morris, R.J., Williams, I.H., Riley, J.R., Smith, A.D., Reynolds, D.R. & Edwards, A.S. (1999). A landscape-scale study of bumble bee foraging range and constancy, using harmonic radar. Journal of Applied Ecology, 36, 519533.Google Scholar
Paine, R.T. (1969). A note on trophic complexity and community stability. The American Naturalist, 103, 9193; doi:10.1086/282586.Google Scholar
Paine, R.T. (1980). Food webs: linkage, interaction strength, and community infrastructure. Journal of Animal Ecology, 49, 667685.Google Scholar
Palka, Z. (1981). On pendant vertices in random graphs. Colloquium Mathematicum, 45, 159167.Google Scholar
Panisson, A., Gauvin, L., Barrat, A. & Cattuto, C. (2012). Fingerprinting temporal networks of close-range human proximity. www.sociopatterns.org.Google Scholar
Pascual, M. & Dunne, J.A. (2006). Ecological Networks: Linking Structure to Dynamics in Food Webs. New York: Oxford University Press.Google Scholar
Pascual-Hortal, L. & Saura, S. (2006). Comparison and development of new graph-based landscape connectivity indices: towards the prioritization of habitat patches and corridors for conservation. Landscape Ecology, 21, 959967.Google Scholar
Paulau, P.V., Feenders, C. & Blasius, B. (2015). Motif analysis in directed ordered networks and applications to food webs. Scientific Reports, 5, 11926; doi:10.1038/srep11926.Google Scholar
Pearl, J. (1988). Probabilistic Reasoning in Intelligent Systems: Networks of Plausible Inference. 2nd ed. San Mateo, CA: Morgan Kaufman.Google Scholar
Pearl, J. (2009). Causality: Models, Reasoning and Inference. 2nd ed. Cambridge. Cambridge University Press.Google Scholar
Penny, D. & Hendy, M.D. (1985). The use of tree comparison metrics. Systematic Zoology, 34, 7582.Google Scholar
Perkins, S.A., Cagnacci, F., Stradiotto, A., Arnoldi, D. & Hudson, P.J. (2009). Comparison of social networks derived from ecological data: implications of inferring infectious disease dynamics. Journal of Animal Ecology, 78, 10151022.Google Scholar
Permogorskiy, M.S. (2015). Competitive intransitivity among species in biotic communities. Biology Bulletin Reviews, 3, 226233.Google Scholar
Petanidou, T., Kallimanis, A.S., Tzanopoulos, J., Sgardelis, S.P. & Pantis, J.D. (2008). Long-term observation of a pollinator network: fluctuation in species and interactions, relative invariance of network structure and implications for estimates of specialization. Ecological Letters, 11, 564575.Google Scholar
Petchey, O.L. & Gaston, K.J. (2002). Functional diversity (FD), species richness, and community composition. Ecology Letters, 5, 402411.Google Scholar
Petchey, O.L. & Gaston, K.J. (2007). Dendrograms and measuring functional diversity. Oikos, 116, 14221426.Google Scholar
Peterson, E.E., Ver Hoef, J.M., Isaak, D.J., Falke, J.A., Fortin, M.-J., Jordan, C., McNyset, K., et al. (2013). Modeling dendritic ecological networks in space: an integrated network perspective. Ecology Letters, 16, 707719.Google Scholar
Peterson, G.D. (2002). Contagious disturbance, ecological memory, and the emergence of landscape pattern. Ecosystems, 5, 329338.Google Scholar
Peterson, G.J., Pressé, S., Peterson, K.S. & Dill, K.A. (2012). Simulated evolution of protein-protein interaction networks with realistic topology. PLoS ONE, 7, e39052; doi:100.1371/journal.pone.0039052.Google Scholar
Petraitis, P.S. (1979). Competitive networks and measures of intransitivity. American Naturalist, 114, 921925.Google Scholar
Peyrard, N., Pellegrin, F., Chadoeuf, J. & Nandris, D. (2006). Statistical analysis of the spatio-temporal dynamics of rubber tree (Hevea brasiliensis) trunk phloem necrosis: no evidence of pathogen transmission. Forest Pathology, 36, 360371.Google Scholar
Phillips, J.D. (2011). The structure of ecological state transitions: amplification, synchronization, and constraints in responses to environmental change. Ecological Complexity, 8, 336346.Google Scholar
Phillips, S.J., Williams, P., Midgley, G. & Archer, A. (2008). Optimizing dispersal corridors for the Cape Proteaceae using network flow. Ecological Applications, 18, 12001211.Google Scholar
Pielou, E.C. (1977). Mathematical Ecology. New York: John Wiley.Google Scholar
Pillai, P., Loreau, M. & Gonzalez, A. (2009). A patch-dynamic framework for food web metacommunities. Theoretical Ecology, 3 doi:10.1007/sl2080-009-0065-1.Google Scholar
Pilosof, S., Fortuna, M.A., Vinarski, M.V., Korallo-Vinarskaya, N.P. & Krasnov, B.R. (2013). Temporal dynamics of direct reciprocal and indirect effects in a host-parasite network. Journal of Animal Ecology, 82, 987996.Google Scholar
Pinkerton, M.H. & Bradford-Grieve, J.M. (2014). Characterizing foodweb structure to identify potential ecosystem effects of fishing in the Ross Sea, Antarctica. ICES Journal of Marine Science, 71 doi:10.1093/icejms/fst230.Google Scholar
Pinto, N. & Keitt, T.H. (2009). Beyond the least-cost path: evaluating corridor redundancy using a graph-theoretic approach. Landscape Ecology, 24, 253266.Google Scholar
Piraveenan, M., Prokopenko, M. & Zomaya, A.Y. (2008). Local assortativeness in scale-free networks. Europhysics Letters, 84, 28002.Google Scholar
Piraveenan, M., Prokopenko, M. & Zomaya, A.Y. (2012). Assortative mixing in directed biological networks. IEEE/ACM Transactions on Computational Biology and Bioinformatics, 9. doi:101109/TCBB.2010.80.Google Scholar
Plavšić, D., Nikolić, S., Trinajstić, N. & Mihalić, Z. (1993). On the Harary Index for the characterization of chemical graphs. Journal of Mathematical Chemistry, 12, 235250.Google Scholar
Podani, J. & Dickinson, T.A. (1984). Comparison of dendrograms: a multivariate approach. Canadian Journal of Botany, 62, 27652778.Google Scholar
Poisot, T., Canard, E., Mouillot, D., Mouquet, N. & Gravel, D. (2012). The dissimilarity of species interaction networks. Ecology Letters, 15, 13531361.Google Scholar
Poisot, T. & Gravel, D. (2014). When is an ecological network complex? Connectance drives degree distribution and emerging network properties. PeerJ, 2, e251; doi 10.7717/peerj.251.Google Scholar
Poisot, T., Stanko, M., Miklisova, D. & Morand, S. (2013). Facultative and obligate parasite communities exhibit different network properties. Parasitology, 140 doi:10.1017/S003118201300851.Google Scholar
Poisot, T., Stouffer, D.B. & Gravel, D. (2015). Beyond species: why ecological interaction networks vary through space and time. Oikos, 124, 243251.Google Scholar
Polis, G.A. & Winemiller, K.O., eds (1996). Food Webs: Integration of Patterns and Dynamics. New York: Chapman & Hall.Google Scholar
Poon, A.F.Y. (2015). Phylodynamic inference with kernel ABC and its application to HIV epidemiology. Molecular Biology and Evolution, 32 doi:10.1093/molbev/msv123.Google Scholar
Poon, A.F.Y., Walker, L.W., Murray, H., McCloskey, R.M., Harrihan, P.R. & Liang, R.H. (2013). Mapping the shapes of phylogenetic trees from human and zoonotic RNA viruses. PLoS ONE, 8, e78122; doi:10.1371/journal.pone.0078122.Google Scholar
Poos, M., Walker, S.C. & Jackson, D.A. (2009). Functional diversity indices can be driven by methodological choices and species richness. Ecology, 90, 341347.Google Scholar
Porphyre, T., Stevenson, M., Jackson, R. & McKenzie, J. (2008). Influence of contact heterogeneity on TB reproduction ratio Ro in a free-living brushtail possum Trchosurus vulpecula population. Veterinary Research, 39, 31.Google Scholar
Poulin, R. (2010). Network analysis shining light on parasite ecology and diversity. Trends in Parasitology, 26, 492498.Google Scholar
Poulin, R. & Guégan, J.-F. (2000). Nestedness, anti-nestedness, and the relationship between prevalence and intensity in ectoparasites assemblages of marine fish: a spatial model of species coexistence. International Journal for Parasitology, 30, 11471152.Google Scholar
Prado, P.I. & Lewinsohn, T.M. (2004). Compartments in insect-plant associations and their consequences for community structure. Journal of Animal Ecology, 73, 11681178.Google Scholar
Proulx, S.R., Promislow, D.E.L. & Phillip, P.C. (2005). Network thinking in ecology and evolution. Trends in Ecology and Evolution, 20, 345352.Google Scholar
Pržulj, K., Wigle, D.A. & Jurisica, I. (2004). Functional topology in a network of protein interactions. Bioinformatics, 20, 340348.Google Scholar
Pržulj, N. (2005). Analyzing Large Biological Networks: Protein-Protein Interactions Example. PhD thesis, Graduate Department of Computer Science, University of Toronto.Google Scholar
Pržulj, N. (2007). Biological network comparison using graphlet degree distribution. Bioinformatics, 23, e177e183.Google Scholar
Pržulj, N. (2010). Protein-protein interactions: making sense of networks via graph-theoretic modelling. Bioessays, 33 doi:10.1002/bies.201000044.Google Scholar
Purchase, H.C. (2002). Metrics for graph drawing aesthetics. Journal of Visual Languages & Computing, 13, 501516.Google Scholar
Raman, K. (2010). Construction and analysis of protein-protein interaction networks. Automated Experimentation, 2, 2; doi:10.1186/1759-4499-2-2.Google Scholar
Raschke, M., Schläpfer, M. & Nibali, R. (2010). Measuring degree-degree association in networks. Physics Review E, 82, 037102.Google Scholar
Rayfield, B., Fortin, M.-J., & Fall, A. (2011). Connectivity for conservation: a framework to classify network measures. Ecology, 92, 847858.Google Scholar
Raymond, B. & Hosie, G. (2009). Network-based exploration and visualization of ecological data. Ecological Modelling, 220, 673683.Google Scholar
Restrepo, G., Mesa, H. & Llanos, E.J. (2007). Three dissimilarity measures to contrast dendrograms. Journal of Chemical Informatics & Modeling, 47, 761770.Google Scholar
Rezende, E.L., Albert, E.M., Fortuna, M.A. & Bascompte, J. (2009). Compartments in a marine food web associated with phylogeny, body mass, and habitat structure. Ecological Letters, 12, 779788.Google Scholar
Richters, O. & Peixoto, T.P. (2011). Trust transitivity in social networks. PLoS ONE, 6, e18384; doi:10.1371/journal.pone.0018384.Google Scholar
Ricotta, C., Stanisci, A., Avena, G. & Blasi, C. (2000). Quantifying the network connectivity of landscape mosaics: a graph-theoretical approach. Community Ecology, 1, 8994.Google Scholar
Robinson, V.B. (2009). Fuzzy sets in spatial analysis. In Fotheringham, A.S. & Rogerson, P.A. (eds), The Sage Handbook of Spatial Analysis, pp. 225241. London: Sage.Google Scholar
Roxburgh, S.H. & Wilson, J.B. (2000). Stability and coexistence in a lawn community: mathematical prediction of stability using a community matrix with parameters derived from competition experiments. Oikos, 88, 395408.Google Scholar
Rozdilsky, I.D., Stone, L. & Solow, A. (2004). The effects of interaction compartments on stability for competitive systems. Journal of Theoretical Biology, 227, 277282.Google Scholar
Rozenfeld, A.F., Arnaud-Haond, S., Hernandez-Garcia, E., Eguiluz, V.M., Serrao, E.A. & Duarte, C.M. (2008). Network analysis identifies weak and strong links in a metapopulation system. Proceedings of the National Academy of Sciences, 105, 1882418829.Google Scholar
Rubinov, M. & Sporns, O. (2010). Complex network measures of brain connectivity: uses and interpretations. NeuroImage, 52, 10591069.Google Scholar
Rubio, L., Bodin, O., Brotons, L. & Saura, S. (2015). Connectivity conservation priorities for individual patches evaluated in the present landscape: how durable and effective are they in the long term? Ecography, 38, 782791.Google Scholar
Ryder, T.B., Horton, B.M., van der Tillaart, M., Morales, J.D.D. & Moore, I.T. (2012). Proximity data-loggers increase the quantity and quality of social network data. Biology Letters, 8 doi:10.1098/rsbl.2012.0536.Google Scholar
Sakr, S. (2013). Processing large-scale graph data: a guide to current technology. IBM developerWorks, /library/os-giraph/.Google Scholar
Salathé, M. & Jones, J.H. (2010). Dynamics and control of diseases in networks with community structure. PLoS Computational Biology, 6, e1000736; doi:10.1371/journal.pcbi.1000736.Google Scholar
Sanfeliu, A. & Fu, K.-S. (1983). A distance measure between attributed relational graphs for pattern recognition. IEEE Transactions on Systems, Man and Cybernetics, 13, 353363; doi:10.1109/TSMC.1983.6313167.Google Scholar
Santos, E. & Young, J.D. (1999). Probabilistic temporal networks: a unified framework for reasoning with time and uncertainty. International Journal of Approximate Reasoning, 20, 263291.Google Scholar
Sarajlić, A., Malod-Dognin, N., Yaveroglu, O.N. & Pržulj, N. (2016). Graphlet-based characterization of directed networks. Scientific Reports, 6, 35098; doi:10.1038/srep35098.Google Scholar
Särkkä, A. & Renshaw, E. (2006). The analysis of marked point patterns evolving through space and time. Computational Statistics and Data Analysis, 51, 16981718.Google Scholar
Satniczenko, P.P.A., Kopp, J.C. & Allesina, S. (2012). The ghost of nestedness in ecological networks. Nature Communications, 4 doi:10.1038/ncomms2422.Google Scholar
Saura, S., Bodin, Ö. & Fortin, M.-J. (2014). Stepping stones are crucial for species’ long-distance dispersal and range expansion through habitat networks. Journal of Applied Ecology, 51, 171182.Google Scholar
Savolainen, R. & Vepsäläinen, K. (1988). A competition hierarchy among boreal ants: impact on resource partitioning and community structure. Oikos, 51, 135155.Google Scholar
Schick, R.S. & Lindley, S.T. (2007). Directed connectivity among fish populations in a riverine network. Journal of Applied Ecology, 44, 11161126.Google Scholar
Schjelderup-Ebbe, T. (1975). Contributions to the social psychology of the domestic chicken. In Schein, M.W. (ed.), Schleidt, M. and Schleidt, W.M. (trans.), Social Hierarchy and Dominance. Benchmark Papers in Animal Behavior, vol. 3, pp. 3549. Stroudsburg, PA: Dowden, Hutchinson and Ross. (Reprinted from Zeitschrift für Psychologie, 1922, 88, 225–252.)Google Scholar
Schoenly, K. & Cohen, J.E. (1991). Temporal variation in food web structure: 16 empirical cases. Ecological Monographs, 61, 267298.Google Scholar
Schölkopf, B., Tsuda, K. & Vert, J.-P. (2004). Kernel Methods on Computational Biology. Cambridge, MA: MIT Press.Google Scholar
Schreiber, S.J. & Killingback, T.P. (2013). Spatial heterogeneity promotes coexistence of rock-paper-scissor metacommunities. Theoretical Population Biology, 86, 111.Google Scholar
Scrosati, R.A. & Heaven, C.S. (2007). Spatial trends in community richness, diversity, and evenness across rocky intertidal environmental stress gradients in Eastern Canada. Marine Ecology Progress Series, 342, 114.Google Scholar
Shawe-Taylor, J. & Cristianini, N. (2004). Kernel Methods for Pattern Analysis. Cambridge: Cambridge University Press.Google Scholar
Shervashidze, N., Schweitzer, P., van Leeuwen, E.J., Mehlhorn, K. & Borgwardt, K.M. (2011). Weisfeiler-Lehman graph kernels. Journal of Machine Learning Research, 12, 25392561.Google Scholar
Shervashidze, N., Vishwanathan, S.V.N., Petri, T., Mehlhorn, K. & Borgwardt, K. (2009). Efficient graphlet kernels for large graph comparison. In Proceedings of the 12th International Conference on Artificial Intelligence and Statistics (AISTATS) 2009, pp. 488495. Clearwater Beach, USA: Society for Artificial Intelligence and StatisticsGoogle Scholar
Shipley, B. (1993). A null model for competitive hierarchies in competition matrices. Ecology, 74, 16931699.Google Scholar
Shipley, B. (2000). Cause and Correlation in Biology. Oxford: Oxford University Press.Google Scholar
Shipley, B. (2009). Confirmatory path analysis in a generalized multilevel context. Ecology, 90, 363368.Google Scholar
Shipley, B. & Keddy, P.A. (1994). Evaluating evidence for competitive hierarchies in plant communities. Oikos, 69, 340345.Google Scholar
Shirley, M.D.F. & Rushton, S.P. (2005). The impacts of network topology on disease spread. Ecological Complexity, 2, 287299.Google Scholar
Shizuki, D. & McDonald, D.B. (2012). A social network perspective on measurements of dominance hierarchies. Faculty Publications in the Biological Sciences, University of Nebraska. http://digitalcommons.unl.edu/bioscifacpub/234.Google Scholar
Siegel, A.F. & Sugihara, G. (1983). Moments of particle size distributions under sequential breakage with applications to species abundance. Journal of Applied Probability, 20, 1516-Aug 4.Google Scholar
Silvertown, J. & Wilson, J.B. (2000). Spatial interactions among grassland plant populations. In Dieckmann, U., Law, R. & Metz, J.A.J. (eds), The Geometry of Ecological Interactions: Simplifying Spatial Complexity, pp. 2847. Cambridge: Cambridge University Press.Google Scholar
Simard, S.W., Beiler, K.J., Bingham, M.A., Deslippe, J.R., Philip, L.J. & Teste, F.P. (2012). Mycorrhizal networks: mechanisms, ecology and modelling. Fungal Biology Reviews, 26, 3960.Google Scholar
Smith, D. (2000). The population dynamics and community ecology of root hemiparasitic plants. American Naturalist, 155, 1323.Google Scholar
Snow, R., Snow, M. & Kufa, N. (2007). GIS analysis of lightning strikes within a tornadic environment. In Proceedings of the 3rd IASME/WSEAS International Conference, Agios, Greece, pp. 466471. WEAS PressGoogle Scholar
Sokal, R.R. & Rohlf, F.J. (1962). The comparison of dendrograms by objective methods. Taxon, 11, 3340.Google Scholar
Solé, R.V. & Montoya, J.M. (2001). Complexity and fragility in ecological networks. Proceedings of the Royal Society London, Series B, 268, 20392045.Google Scholar
Soliveres, S., et al. (2015). Intransitive competition is widespread in plant communities and maintains their species richness. Ecology Letters, 18, 790798; doi:10.1111/ele.12456.Google Scholar
Sotomayor, D.A. & Lortie, C.J. (2015). Indirect interactions in terrestrial plant communities: emerging patterns and research gaps. Ecosphere, 6, 103; doi:10.1890/ES14-00117.1.Google Scholar
Soula, S. & Chauzy, S. (2001). Some aspects of the correlation between lightning and rain activity in thunderstorms. Atmospheric Research, 56, 355373.Google Scholar
Southworth, D., He, X.-H., Swenson, W., Bledsoe, C.S. & Horwath, W.R. (2005). Application of network theory to potential mycorrhizal networks. Mycorrhiza, 15, 589595.Google Scholar
Spiesman, B.J. & Inouye, B.D. (2013). Habitat loss alters the architecture of plant-pollinator interaction networks. Ecology, 94, 26882696.Google Scholar
Sporns, O., Honey, C.J. & Kötter, R. (2007). Identification and classification of hubs in brain networks. PLoS ONE, 2 (10), e1049; doi:01.1371/journal.pone.0001049.Google Scholar
Sporns, O. & Kötter, R. (2004). Motifs in brain networks. PLoS Biology, 2. doi:10.1371/journal.pbio.0020369.Google Scholar
Staniczenko, P.P.A., Kopp, J.C. & Allesina, S. (2012). The ghost of nestedness in ecological networks. Nature Communications, 4, 1391; doi:10.1038/ncomms2422.Google Scholar
Starling, F.L. do R.M. 2000. Comparative study of the zooplankton composition of six lacustrine ecosystems in central Brazil during the dry season. Review of Brasilian Biology, 60, 101111.Google Scholar
Stouffer, D.B. & Bascompte, J. (2010). Understanding food-web persistence from local to global scales. Ecology Letters, 13, 154161.Google Scholar
Stouffer, D.B. & Bascompte, J. (2011). Compartmentalization increases food web persistence. Proceedings of the National Academy of Sciences, 108, 36483652.Google Scholar
Stouffer, D.B., Camacho, J. & Amaral, L.A.N. (2006). A robust measure of food web intervality. Proceedings of the National Academy of Sciences, 103, 1901519020.Google Scholar
Stouffer, D.B., Camacho, J., Jiang, W. & Amaral, L.A.N. (2007). Evidence for the existence of a robust pattern of prey selection in food webs. Proceedings of the Royal Society, 274, 19311940.Google Scholar
Stoyan, D., Kendall, W.S. & Mecke, J. (1995). Stochastic Geometry and Its Applications. 2nd ed. Chichester: Wiley.Google Scholar
Strauss, R.E. (1982). Statistical significance of species clusters in association analysis. Ecology, 63, 634639.Google Scholar
Strickland, C., Dangelmayr, G., Shipman, P.D., Kumar, S. & Stohlgren, T.J. (2015). Network spread of invasive species and infectious diseases. Ecological Modelling, 309–310, 19.Google Scholar
Strimmer, K. & Moulton, V. (2000). Likelihood analysis of phylogenetic networks using directed graphical models. Molecular Biology & Evolution, 17, 875881.Google Scholar
Sugihara, G. (1980). Minimal community structure: an explanation of species abundance patterns. American Naturalist, 116, 770787.Google Scholar
Sugihara, G. (1982). Niche Hierarchy: Structure, Organization, and Assembly in Natural Communities. PhD dissertation, Princeton University.Google Scholar
Sugihara, G. (1984). Graph theory, homology and food webs. In Levin, S.A. (ed.), Population Biology: Proceedings of Symposia in Applied Mathematics, pp. 83101. Providence, RI: American Mathematical Society.Google Scholar
Sugihara, G., Bersier, L.-F., Southwood, T.R.E., Pimm, S.L. & May, R.M. (2003). Predicted correspondence between species abundances and dendrograms of niche similarities. Proceedings of the National Academy of Sciences, 100, 52465251.Google Scholar
Sun, J., Zhang, M. & Tan, C.-L. (2011). Tree sequence kernel for natural language. In Proceedings of the Twenty-Fifth AAAI Conference on Artificial Intelligence, pp. 921926. Menlo Park, USA: AAAI PressGoogle Scholar
Suweis, S., Simini, F., Banavar, J.R. & Maritan, A. (2013). Emergence of structural and dynamical properties of ecological mutualistic networks. Nature, 500, 449452.Google Scholar
Takahashi, K., Kulldorff, M., Tango, T. & Yih, K. (2008). A flexibly shaped space-time scan statistic for disease outbreak detection and monitoring. International Journal of Health Geographics, 7, 14; doi:10.1186/1476-072X-7-14.Google Scholar
Tang, J., Leontiadis, I., Scellato, S., Nicosia, V., Mascolo, C., Musolesi, M. & Latora, V. (2013). Applications of temporal graph metrics to real-world networks. In Holme, P. & Saramäki, J. (eds), Temporal Networks, pp. 135159. Berlin: Springer.Google Scholar
Tang, J., Musolesi, M., Mascolo, C. & Latora, V. (2009). Temporal distance metrics for social network analysis. Paper presented at the 2nd SIGCOMM Workshop on Online Social Networks, Barcelona.Google Scholar
Tango, T. (2010). Statistical Methods for Disease Clustering. New York: Springer.Google Scholar
Tanner, J.E., Hughes, T.P. & Connell, J.H. (1996). The role of history in community dynamics: a modeling approach. Ecology, 77, 108117.Google Scholar
Teng, J. & McCann, K.S. (2004). Dynamics of compartmented and reticulate food webs in relation to energetic flows. American Naturalist, 164, 85100.Google Scholar
Tero, A., Takagi, S., Saigusa, T., Ito, K., Bebber, D.P., Fricker, M.D., Yumiki, K., Kobayashi, R. & Nakagi, T. (2010). Rules for biologically inspired adaptive network design. Science, 327, 439442.Google Scholar
Thébault, E. & Fontaine, C. (2010). Stability of ecological communities and the architecture of mutualistic and trophic networks. Science, 329, 853856; doi:10.1126/science.1188321.Google Scholar
Thedchanamoorthy, G., Piraveenan, M., Kasthuriratne, D. & Senanayake, U. (2014). Node assortativity in complex networks: an alternative approach. Procedia Computer Science, 29, 24 Feb-49461; doi:10.1016/j.procs.2014.05.229.Google Scholar
Theobald, D.M. (2006). Exploring the functional connectivity of landcapes using landscape networks. In Crooks, K.R. & Sanjayan, M. (eds), Connectivity Conservation, pp. 416444. Cambridge: Cambridge University Press.Google Scholar
Thibaud, R., Del Mondo, G., Garlan, T., Mascret, A. & Carpentier, C. (2013). A spatio-temporal graph model for marine dune dynamics analysis and representation. Transactions in GIS, 17, 742762.Google Scholar
Thomas, A.G. & Dale, M.R.T. (1991). Weed communities in spring-seeded crops in Manitoba. Canadian Journal of Plant Science, 71, 10691080.Google Scholar
Thompson, J.N. (2006). Mutualistic webs of species. Science, 312, 372373.Google Scholar
Thompson, R.M., Hemberg, M., Starzomski, B.M. & Shurin, J.B. (2007). Trophic levels and trophic tangles: the prevalence of omnivory in real food webs. Ecology, 88, 612617.Google Scholar
Thornton, B.M., Knowlton, J.L. & Kuntz, W.A. (2015). Interspecific competition and social hierarchies in frugivorous birds of Costa Rica. Journal of Young Investigators, 29 (2), 15.Google Scholar
Toju, H., Guimaräes, P.R., Olesan, J.M. & Thompson, J.N. (2015). Below-ground plant-fungus network topology is not congruent with above-ground plant-animal network topology. Science Advances, 1 doi:10.1126/sciadv.1500291.Google Scholar
Tosa, M.I., Schauber, E.M. & Nielsen, C.K. (2013). Familiarity breeds contempt: combining proximity loggers and GPS reveals female white-tailed deer (Odocoileus virginianus) avoiding close contact with neighbours. Journal of Wildlife Diseases, 51, 7088.Google Scholar
Toussaint, G.T. (1980). The relative neighbourhood graph of a finite planar set. Pattern Recognition, 12, 261268.Google Scholar
Treml, E.A., Halpin, P.N., Urban, D.L. & Pratson, L.F. (2008). Modeling population connectivity by ocean currents, a graph-theoretic approach for marine conservation. Landscape Ecology, 23, 1936.Google Scholar
Trpevski, I., Dimitrova, T., Boshkovski, T., Stikov, N. & Kocarev, L. (2016). Graphlet characteristics in directed networks. Scientific Reports, 6, 37057; doi:10.1038/srep37057.Google Scholar
Tufte, E.R. (1983). The Visual Display of Quantitative Information. Cheshire, CT: Graphics Press.Google Scholar
Tutte, W.T. (1998). Graph Theory as I Have Known It. Oxford: Clarendon Press.Google Scholar
Ulanowicz, R.E. & Puccia, C.J. (1990). Mixed trophic impacts in ecosystems. Coenoses, 5, 716.Google Scholar
Ulrich, W. (2009). Ecological interaction networks: prospects and pitfalls. Ecological Questions, 11, 1725.Google Scholar
Ulrich, W., Almeida-Neto, M. & Gotelli, N.J. (2009). A consumer's guide to nestedness analysis. Oikos, 118, 317.Google Scholar
Ulrich, W., Soliveres, S., Kryszwski, W., Maestre, F.T. & Gotelli, N.J. (2014). Matrix models for quantifying competitive intransitivity from species abundance data. Oikos, 123, 10571070; doi:10.1111/oik.01217.Google Scholar
Urban, D.L. & Keitt, T. (2001). Landscape connectivity: a graph-theoretic perspective. Ecology, 82, 12051218.Google Scholar
Urban, D.L., Minor, E.S., Treml, E.A. & Schick, R.S. (2009). Graph models of habitat mosaics. Ecology Letters, 12, 260273.Google Scholar
Valls, A., Coll, M. & Christensen, V. (2015). Keystone species: toward an operational concept for marine biodiversity conservation. Ecological Monographs, 85, 2947.Google Scholar
van Langevelde, F., van der Knaap, W.G.M. & Claassen, G.D.H. (1998). Comparing connectivity in landscape networks. Environment and Planning B: Planning and Design, 25, 849863.Google Scholar
van Veen, F.J.F., Morris, R.J. & Godfray, H.C.J. (2006). Apparent competition, quantitative food webs, and the structure of phytophagous insect communities. Annual Review of Entomology, 51, 187208.Google Scholar
van Wijk, B.C.M., Stam, C.J. & Daffertshofer, A. (2010). Comparing brain networks of different size and connectivity density using graph theory. PLoS ONE, 5 (10), e13701.Google Scholar
Vázquez, A. (2013). Spreading dynamics following bursty activity patterns. In Holme, P. & Saramäki, J. (eds), Temporal Networks, pp. 161173. Berlin: Springer.Google Scholar
Vázquez, A., Flammini, A., Maritan, A. & Vespignani, A. (2003). Global protein function prediction from protein-protein interaction networks. Nature Biotechnology, 21, 697700; doi:10/1038/nbt825.Google Scholar
Vázquez, D.P., Chacoff, N.P. & Cagnolo, L. (2009). Evaluating multiple determinants of the structure of plant-animal mutualistic networks. Ecology, 90, 20392046.Google Scholar
Vázquez, D.P., Morris, W.F. & Jordano, P. (2005a). Interaction frequency as a surrogate for the total effect of animal mutualists on plants. Ecology Letters, 8, 10881094.Google Scholar
Vázquez, D.P., Poulin, R., Krasnov, B.R. & Shenbrot, G.I. (2005b). Species abundance and the distribution of specialization in host-parasite interaction networks. Journal of Animal Ecology, 74, 946955.Google Scholar
Vehrencamp, S.L. (1983). A model for the evolution of despotic versus egalitarian societies. Animal Behaviour, 31, 667682.Google Scholar
Vepsäläinen, K. & Czechowski, W. (2014). Against the odds of the ant competition hierarchy: submissive Myrmica rugulosa block access of the dominant Lasius fuliginosus to its aphids. Insectes Sociaux, 61, 8993.Google Scholar
Ver Hoef, J., Peterson, E. & Theobald, D. (2006). Spatial statistical models that use flow and stream distance. Environmental and Ecological Statistics, 13, 449464.Google Scholar
Vinayagam, A., Zirin, J., Roesel, C., Hu, Y., Yilmazel, B., Samsonova, A.A., Neumüller, R., Mohr, S.E. & Perrimon, N. (2014). Integrating protein-protein interaction networks with phenotypes reveals signs of interactions. Nature Methods, 11, 9499; doi:10.1038/nmeth.2733.Google Scholar
Vishwanathan, S.V.N., Schraudolph, N.N., Kondor, R. & Borgwardt, K.M. (2010). Graph kernels. Journal of Machine Learning Research, 11, 12011242.Google Scholar
Wagner, H.H. & Fortin, M.-J. (2013). A conceptual framework for the spatial analysis of landscape genetic data. Conservation Genetics, 14, 253261.Google Scholar
Walker, R., Arima, E., Messina, J., Soares-Filho, B., Perz, S., Vergara, D., Sales, M., Pereira, R. & Castro, W. (2013). Modeling spatial decisions with graph theory: logging roads and forest fragmentation in the Brazilian Amazon. Ecological Applications, 23, 239254.Google Scholar
Walker, S.C., Poos, M.S. & Jackson, D.A. (2008). Functional rarefaction: estimating functional diversity from field data. Oikos, 117, 286296.Google Scholar
Wang, Y.-H., Yang, K.-C., Bridgman, C.L. & Lin, L.-K. (2008). Habitat suitability modeling to correlate gene flow with landscape connectivity. Landscape Ecology, 23, 9891000.Google Scholar
Wardle, G.M. (1998). A graph theory approach to demographic loop analysis. Ecology, 79, 25392549.Google Scholar
Waser, N.M. & Ollerton, J. (2006). Plant-Pollinator Interactions. Chicago: University of Chicago Press.Google Scholar
Wasserman, S. & Faust, K. (1994). Social Network Analysis: Methods and Applications. Cambridge: Cambridge University Press.Google Scholar
Watts, A.G., Schlichting, P., Billerman, S., Jesmer, B., Micheletti, S., Fortin, M.-J., Funk, W.C., Hapeman, P., Muths, E. & Murphy, M.A. (2015). How spatio-temporal habitat connectivity affects amphibian genetic structure. Frontiers in Genetics, 6, Article 275.Google Scholar
Watts, D.J. (1999). Small Worlds: The Dynamics of Networks between Order and Randomness. Princeton, NJ: Princeton University Press.Google Scholar
Watts, D.J. (2003). Six Degrees: The Science of a Connected Age. New York: W.W. Norton.Google Scholar
Webb, C.O., Ackerly, D.D., McPeek, M.A. & Donaghue, M.J. (2002). Phylogenies and community ecology. Annual Review of Ecology and Systematics, 33, 475505.Google Scholar
Weigelt, A., Schumacher, J., Walther, T., Bartelheimer, M., Steinlein, T. & Beyschlag, W. (2007). Identifying mechanisms of competition in multi-species communities. Journal of Ecology, 95, 5364.Google Scholar
West, D.B. (2001). Introduction to Graph Theory. 2nd ed. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Wey, T., Blumstein, D.T., Shen, W. & Jordán, F. (2008). Social network analysis of animal behavior: a promising tool for the study of sociality. Animal Behaviour, 75, 333344.Google Scholar
Wilbur, H.M. (1997). Experimental ecology of foodwebs: complex systems in temporary ponds. Ecology, 78, 22792302.Google Scholar
Williams, M.J. & Musolesi, M. (2016). Spatio-temporal networks: reachability, centrality and robustness. Royal Society Open Science, 3 doi:10.1098/rsos.160196.Google Scholar
Williams, R.J. & Martinez, N.D. (2004). Limits to trophic levels and omnivory in complex food webs: theory and data. American Naturalist, 163, 458468.Google Scholar
Williams, R.J., Berlow, E.L., Dunne, J.A. & Martinez, N.D. (2002). Two degrees of separation in complex food webs. Proceedings of the National Academy of Sciences, 99, 1291312916.Google Scholar
Winemiller, K.O. (1990). Spatial and temporal variation in tropical fish trophic networks. Ecological Monographs, 60, 331367.Google Scholar
Winterbach, W., Van Mieghem, P., Reinders, M.J.T., Wang, H. & de Ridder, D. (2013). Local topological signatures for network-based prediction of biological function. In Pattern Recognition in Bioinformatics, Proceedings of the 8th IAPR conference, Nice 2013, pp. 2324. Berlin: Springer.Google Scholar
Wolfram, S. (2002). A New Kind of Science. Champagne: Wolfram Media.Google Scholar
Woodward, G., Ebenman, B., Emmeson, M., Montoya, J.M., Olesen, J., Valido, A. & Warren, P.H. (2005). Body size in ecological networks. Trends in Ecology and Evolution, 20. doi:10.116/j.tree.2005.04.05.Google Scholar
Woodward, G., et al. (2010). Ecological networks in a changing climate. Advances in Ecological Research, 42, 72122; doi:10.1016/S0065-2504(10)42002-4.Google Scholar
Wu, H., Cheng, J., Huang, S., Ke, Y. & Xu, Y. (2014). Path problems in temporal graphs. Proceedings of the VLDB Endowment, 7, 721729.Google Scholar
Yaveroglu, O.N., Fitzhugh, S.M., Kurant, M., Markopoulou, A., Butts, C.T. & Pržulj, N. (2015). ergm.graphlets: a package for ERG modeling based on graphlet statistics. Journal of Statistical Software, 65 (12), 129Google Scholar
Yaveroglu, O.N., Malod-Dognin, N., Davis, D., Levnajic, Z., Janjic, V., Karapandza, R., Stojmirovic, A. & Pržulj, N. (2014). Revealing the hidden language of complex networks. Scientific Reports, 4, 4547; doi:10.1038/srep04547 (2014).Google Scholar
Yazdani, A. & Jeffrey, P. (2010). A complex network approach to robustness and vulnerability of spatially organized distribution networks. arXiv:1008.1770v2.Google Scholar
Yu, D., Liu, Y., Xun, B. & Shao, H. (2013). Measuring landscape connectivity in an urban area for biological conservation. Clean – Soil, Air, Water, 41, 19.Google Scholar
Yugandhar, K. & Gromiha, M.M. (2016). Analysis of protein-protein interaction networks based on binding affinity. Current Protein and Peptide Science, 17, 7281.Google Scholar
Zhao, K., Karsai, M. & Bianconi, G. (2013a). Models, entropy and information of temporal social networks. In Holme, P. & Saramäki, J. (eds), Temporal Networks, pp. 95117. Berlin: Springer.Google Scholar
Zhao, Q., Azeria, E.T., Le Blanc, M.-L., Lemaître, J. & Fortin, D. (2013b). Landscape-scale disturbances modified bird community dynamics in successional forest environment. PLoS ONE, 8, e81358; doi:10.1371/journal.pone.0081358.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Mark R.T. Dale, University of Northern British Columbia
  • Book: Applying Graph Theory in Ecological Research
  • Online publication: 31 October 2017
  • Chapter DOI: https://doi.org/10.1017/9781316105450.015
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Mark R.T. Dale, University of Northern British Columbia
  • Book: Applying Graph Theory in Ecological Research
  • Online publication: 31 October 2017
  • Chapter DOI: https://doi.org/10.1017/9781316105450.015
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Mark R.T. Dale, University of Northern British Columbia
  • Book: Applying Graph Theory in Ecological Research
  • Online publication: 31 October 2017
  • Chapter DOI: https://doi.org/10.1017/9781316105450.015
Available formats
×