Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-c47g7 Total loading time: 0 Render date: 2024-04-25T02:15:12.591Z Has data issue: false hasContentIssue false

15 - Trait-Based and Process-Oriented Modeling in Ecological Network Dynamics

from Part II - Food Webs: From Traits to Ecosystem Functioning

Published online by Cambridge University Press:  05 December 2017

John C. Moore
Affiliation:
Colorado State University
Peter C. de Ruiter
Affiliation:
Wageningen Universiteit, The Netherlands
Kevin S. McCann
Affiliation:
University of Guelph, Ontario
Volkmar Wolters
Affiliation:
Justus-Liebig-Universität Giessen, Germany
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
Adaptive Food Webs
Stability and Transitions of Real and Model Ecosystems
, pp. 228 - 256
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

Allesina, S. and Bodini, A. (2004). Who dominates whom in the ecosystem? Energy flow bottlenecks and cascading extinctions. Journal of Theoretical Biology, 230, 351358.Google Scholar
Allesina, S., Bodini, A., and Bondavalli, C. (2005). Ecological subsystems via graph theory: the role of strongly connected components. Oikos, 110, 164176.Google Scholar
Barton, R. A., Byrne, R. W., and Whiten, A. (1996). Ecology, feeding competition and social structure in baboons. Behavioral Ecology and Sociobiology, 38, 321329.Google Scholar
Berlow, E. L., Neutel, A. M., Cohen, J. E., et al. (2004). Interaction strengths in food webs: issues and opportunities. Journal of Animal Ecology, 73, 585598.Google Scholar
Berlow, E. L., Dunne, J. A., Martinez, N. D., et al. (2009). Simple prediction of interaction strengths in complex food webs. Proceedings of the National Academy of Sciences of the United States of America, 106, 187191.Google Scholar
Black, A. J. and McKane, A. J. (2012). Stochastic formulation of ecological models and their applications. Trends in Ecology and Evolution, 27, 337345.CrossRefGoogle ScholarPubMed
Bolnick, D. I., Svanbäck, R., Araújo, M. S., and Persson, L. (2007). Comparative support for the niche variation hypothesis that more generalized populations also are more heterogeneous. Proceedings of the National Academy of Sciences of the United States of America, 104, 1007510079.Google Scholar
Bolnick, D. I., Amarasekare, P., and Araújo, M. S. (2011). Why intraspecific trait variation matters in community ecology. Trends in Ecology and Evolution, 26, 183192.CrossRefGoogle ScholarPubMed
Bondavalli, C. and Ulanowicz, R. E. (1999). Unexpected effects of predators upon their prey: the case of the American alligator. Ecosystems, 2, 4963.Google Scholar
Botkin, D. B., Janak, J. F., and Wallis, J. R. (1972). Some ecological consequences of a computer model of forest growth. Journal of Ecology, 60, 849872.CrossRefGoogle Scholar
Brose, U. (2008). Complex food webs prevent competitive exclusion among producer species. Proceedings of the Royal Society B: Biological Sciences, 275, 25072514.CrossRefGoogle ScholarPubMed
Carnicer, J., Brotons, L., Stefanescu, C., and Penuelas, J. (2012). Biogeography of species richness gradients: linking adaptive traits, demography and diversification. Biological Reviews, 87, 457479.CrossRefGoogle ScholarPubMed
Crooks, K. R. and Soulé, M. E. (1999). Mesopredator release and avifaunal extinctions in a fragmented system. Nature, 400, 563566.CrossRefGoogle Scholar
DeAngelis, D. L. and Gross, L. J. (1992). Individual-Based Models and Approaches in Ecology: Populations, Communities and Ecosystems. New York: Chapman and Hall.Google Scholar
DeAngelis, D. L., Cox, D. K., and Coutant, C. C. (1980). Cannibalism and size dispersal in young-of-the-year largemouth bass: experiment and model. Ecological Modelling, 8, 133148.CrossRefGoogle Scholar
DeAngelis, D. L., Loftus, W. F., Trexler, J. C., and Ulanowicz, R. E. (1997). Modeling fish dynamics and effects of stress in a hydrologically pulsed ecosystem. Journal of Aquatic Ecosystem Stress and Recovery, 6, 113.Google Scholar
Devijver, P. A. and Kittler, J. (1982). Pattern Recognition: A Statistical Approach (Vol. 761). London: Prentice-Hall.Google Scholar
dit Durell, S. E., Stillman, R. A., Caldow, R. W., et al. (2006). Modelling the effect of environmental change on shorebirds: a case study on Poole Harbour, UK. Biological Conservation, 131, 459473.CrossRefGoogle Scholar
Eklöf, A., Jacob, U., Kopp, J., et al. (2013). The dimensionality of ecological networks. Ecology Letters, 16, 577583.Google Scholar
Froese, R. and Pauly, D. (eds.) (2000). FishBase 2000: Concepts, Design and Data Sources (No. 1594). WorldFish.Google Scholar
Gamarra, J. G. and Solé, R. V. (2002). Complex discrete dynamics from simple continuous population models. Bulletin of Mathematical Biology, 64, 611620.Google Scholar
Gergs, A. and Ratte, H. T. (2009). Predicting functional response and size selectivity of juvenile Notonecta maculata foraging on Daphnia magna. Ecological Modelling, 220, 33313341.Google Scholar
Giacomini, H. C., De Marco, P. Jr., and Petrere, M. Jr. (2009). Exploring community assembly through an individual based model for trophic interactions. Ecological Modelling, 220, 2339.Google Scholar
Giacomini, H. C., DeAngelis, D. L., Trexler, J. C., and Petrere, M. Jr. (2013). Trait contributions to fish community assembly emerge from trophic interactions in an individual-based model. Ecological Modelling, 251, 3243.Google Scholar
Gjata, N., Scotti, M., and Jordán, F. (2012). The strength of simulated indirect interaction modules in a real food web. Ecological Complexity, 11, 160164.Google Scholar
Grimm, V. (1999). Ten years of individual-based modelling in ecology: what have we learned and what could we learn in the future? Ecological Modelling, 115, 129148.Google Scholar
Grimm, V. and Railsback, S. F. (2013). Individual-Based Modeling and Ecology. Princeton University Press.Google Scholar
Grimm, V., Revilla, E., Berger, U., et al. (2005). Pattern-oriented modeling of agent-based complex systems: lessons from ecology. Science, 310, 987991.Google Scholar
Grimm, V., Berger, U., Bastiansen, F., et al. (2006). A standard protocol for describing individual-based and agent-based models. Ecological Modelling, 198, 115126.Google Scholar
Grimm, V., Berger, U., DeAngelis, D. L., et al. (2010). The ODD protocol: a review and first update. Ecological Modelling, 221, 27602768.Google Scholar
Gross, T., Rudolf, L., Levin, S. A., and Dieckmann, U. (2009). Generalized models reveal stabilizing factors in food webs. Science, 325, 747750.Google Scholar
Hanski, I., Pakkala, T., Kuussaari, M., and Lei, G. (1995). Metapopulation persistence of an endangered butterfly in a fragmented landscape. Oikos, 72, 2128.CrossRefGoogle Scholar
Hartig, F., Calabrese, J. M., Reineking, B., Wiegand, T., and Huth, A. (2011). Statistical inference for stochastic simulation models: theory and application. Ecology Letters, 14, 816827.Google Scholar
Hartvig, M. and Andersen, K. H. (2013). Coexistence of structured populations with size-based prey selection. Theoretical Population Biology, 89, 2433.Google Scholar
Hebblewhite, M. and Merrill, E. H. (2007). Multiscale wolf predation risk for elk: does migration reduce risk? Oecologia, 152, 377387.Google Scholar
Huse, G., Johansen, G. O., Bogstad, B., and Gjøsæter, H. (2004). Studying spatial and trophic interactions between capelin and cod using individual-based modelling. ICES Journal of Marine Science, 61, 12011213.CrossRefGoogle Scholar
Ibarrola, I., Arambalza, U., Navarro, J. M., Urrutia, M. B., and Navarro, E. (2012). Allometric relationships in feeding and digestion in the Chilean mytilids Mytilus chilensis (Hupé), Choromytilus chorus (Molina) and Aulacomya ater (Molina): a comparative study. Journal of Experimental Marine Biology and Ecology, 426–427, 1827.Google Scholar
Jeltsch, F., Müller, M. S., Grimm, V., Wissel, C., and Brandl, R. (1997). Pattern formation triggered by rare events: lessons from the spread of rabies. Proceedings of the Royal Society B: Biological Sciences, 264, 495503.CrossRefGoogle ScholarPubMed
Jones, A. G., Arnold, S. J., and Bürger, R. (2003). Stability of the G‐matrix in a population experiencing pleiotropic mutation, stabilizing selection, and genetic drift. Evolution, 57, 17471760.Google Scholar
Jordán, F., Scotti, M., and Priami, C. (2011). Process algebra-based computational tools in ecological modelling. Ecological Complexity, 8, 357363.Google Scholar
Kéfi, S., Berlow, E. L., Wieters, E. A., et al. (2012). More than a meal… integrating non‐feeding interactions into food webs. Ecology Letters, 15, 291300.Google Scholar
Kunz, H. and Hemelrijk, C. K. (2003). Artificial fish schools: collective effects of school size, body size, and body form. Artificial Life, 9, 237253.Google Scholar
Livi, C. M., Jordán, F., Lecca, P., and Okey, T. A. (2011). Identifying key species in ecosystems with stochastic sensitivity analysis. Ecological Modelling, 222, 25422551.Google Scholar
Lorek, H. and Sonnenschein, M. (1999). Modelling and simulation software to support individual-based ecological modelling. Ecological Modelling, 115, 199216.CrossRefGoogle Scholar
McCann, K. S. (2000). The diversity–stability debate. Nature, 405, 228233.Google Scholar
McGill, B. J., Enquist, B. J., Weiher, E., and Westoby, M. (2006). Rebuilding community ecology from functional traits. Trends in Ecology and Evolution, 21, 178185.Google Scholar
Megrey, B. A. and Hinckley, S. (2001). Effect of turbulence on feeding of larval fishes: a sensitivity analysis using an individual-based model. ICES Journal of Marine Science, 58, 10151029.Google Scholar
Melián, C. J., Vilas, C., Baldó, F., et al. (2011). Eco-evolutionary dynamics of individual-based food webs. Advances in Ecological Research, 45, 226.Google Scholar
Meyer, K. M., Wiegand, K., Ward, D., and Moustakas, A. (2007). The rhythm of savanna patch dynamics. Journal of Ecology, 95, 13061315.Google Scholar
Meyer, K. M., Mooij, W. M., Vos, M., Hol, W. H. G., and van der Putten, W. H. (2009). The power of simulating experiments. Ecological Modelling, 220, 25942597.Google Scholar
Moore, J. C., Boone, R. B., Koyama, A., and Holfelder, K. (2014). Enzymatic and detrital influences on the structure, function, and dynamics of spatially-explicit model ecosystems. Biogeochemistry, 117, 205227.Google Scholar
Morita, K. and Yokota, A. (2002). Population viability of stream-resident salmonids after habitat fragmentation: a case study with white-spotted charr Salvelinus leucomaenis by an individual based model. Ecological Modelling, 155, 8594.Google Scholar
Neutel, A. M., Heesterbeek, J. A., and de Ruiter, P. C. (2002). Stability in real food webs: weak links in long loops. Science, 296, 11201123.CrossRefGoogle ScholarPubMed
Newman, T. J., Ferdy, J. B., and Quince, C. (2004). Extinction times and moment closure in the stochastic logistic process. Theoretical Population Biology, 65, 115126.Google Scholar
Odum, H. T. (1956). Efficiencies, size of organisms, and community structure. Ecology, 37, 592597.Google Scholar
Otto, S. B., Rall, B. C., and Brose, U. (2007). Allometric degree distributions facilitate food-web stability. Nature, 450, 12261229.Google Scholar
Parry, H. R., Topping, C. J., Kennedy, M. C., Boatman, N. D., and Murray, A. W. (2013). A Bayesian sensitivity analysis applied to an agent-based model of bird population response to landscape change. Environmental Modelling and Software, 45, 104115.Google Scholar
Peters, R. H. (1993). The Ecological Implications of Body Size. Cambridge University Press.Google Scholar
Pimm, S. L. (1982). Food Webs. London: Chapman and Hall.Google Scholar
Pitt, W. C., Box, P. W., and Knowlton, F. F. (2003). An individual-based model of canid populations: modelling territoriality and social structure. Ecological Modelling, 166, 109121.Google Scholar
Pratchett, M. S., Wilson, S. K., Graham, N. A. J., et al. (2009). Coral bleaching and consequences for motile reef organisms: past, present and uncertain future effects. In Coral Bleaching, eds. van Oppen, M. J. H. and Lough, J. M., Berlin Heidelberg: Springer, pp. 139158.CrossRefGoogle Scholar
Pugnaire, F. I., Haase, P., and Puigdefábregas, J. (1996). Facilitation between higher plant species in a semiarid environment. Ecology, 77, 14201426.Google Scholar
Rademacher, C., Neuert, C., Grundmann, V., Wissel, C., and Grimm, V. (2004). Reconstructing spatiotemporal dynamics of Central European natural beech forests: the rule-based forest model BEFORE. Forest Ecology and Management, 194, 349368.CrossRefGoogle Scholar
Railsback, S. F. and Harvey, B. C. (2002). Analysis of habitat-selection rules using an individual-based model. Ecology, 83, 18171830.Google Scholar
Rall, B. C., Brose, U., Hartvig, M., et al. (2012). Universal temperature and body-mass scaling of feeding rates. Philosophical Transactions of the Royal Society B: Biological Sciences, 367, 29232934.Google Scholar
Reynolds, C. W. (1987). Flocks, herds and schools: a distributed behavioral model. Computer Graphics, 21, 2534.CrossRefGoogle Scholar
Rose, K. A., Rutherford, E. S., McDermot, D. S., Forney, J. L., and Mills, E. L. (1999). Individual-based model of yellow perch and walleye populations in Oneida Lake. Ecological Monographs, 69, 127154.Google Scholar
Ruel, J. J. and Ayres, M. P. (1999). Jensen’s inequality predicts effects of environmental variation. Trends in Ecology and Evolution, 14, 361366.Google Scholar
Ruesink, J. L. (1998). Variation in per capita interaction strength: thresholds due to nonlinear dynamics and nonequilibrium conditions. Proceedings of the National Academy of Sciences of the United States of America, 95, 68436847.Google Scholar
Saeys, Y., Inza, I., and Larrañaga, P. (2007). A review of feature selection techniques in bioinformatics. Bioinformatics, 23, 25072517.Google Scholar
Scotti, M., Gjata, N., Livi, C. M., and Jordán, F. (2012). Dynamical effects of weak trophic interactions in a stochastic food web simulation. Community Ecology, 13, 230237.Google Scholar
Scotti, M., Ciocchetta, F., and Jordán, F. (2013). Social and landscape effects on food webs: a multi-level network simulation model. Journal of Complex Networks, 1, 160182.CrossRefGoogle Scholar
Shin, Y. J. and Cury, P. (2001). Exploring fish community dynamics through size-dependent trophic interactions using a spatialized individual-based model. Aquatic Living Resources, 14, 6580.Google Scholar
Shugart, H. H. (1984). A Theory of Forest Dynamics: The Ecological Implications of Forest Succession Models. New York: Springer-Verlag.Google Scholar
Souissi, S., Seuront, L., Schmitt, F. G., and Ginot, V. (2005). Describing space-time patterns in aquatic ecology using IBMs and scaling and multi-scaling approaches. Nonlinear Analysis: Real World Applications, 6, 705730.Google Scholar
Stouffer, D. B. and Bascompte, J. (2011). Compartmentalization increases food-web persistence. Proceedings of the National Academy of Sciences of the United States of America, 108, 36483652.Google Scholar
Straile, D. (2002). North Atlantic Oscillation synchronizes food-web interactions in central European lakes. Proceedings of the Royal Society B: Biological Sciences, 269, 391395.Google Scholar
Tofts, C. (1993). Algorithms for task allocation in ants. (A study of temporal polyethism: theory.) Bulletin of Mathematical Biology, 55, 891918.Google Scholar
van Nes, E. H., Lammens, E. H., and Scheffer, M. (2002). PISCATOR, an individual-based model to analyze the dynamics of lake fish communities. Ecological Modelling, 152, 261278.Google Scholar
Williams, R. J. and Martinez, N. D. (2004). Stabilization of chaotic and non-permanent food-web dynamics. European Physical Journal B-Condensed Matter and Complex Systems, 38, 297303.Google Scholar
Winemiller, K. O. (2017). Food web dynamics when divergent life history strategies respond to environmental variation differently: a fisheries ecology perspective. In Adaptive Food Webs: Stability and Transitions of Real and Model Ecosystems, ed. de Ruiter, P. C., Wolters, V., McCann, K. S., and Moore, J. C.. Cambridge: Cambridge University Press, pp. 305323.Google Scholar
Yamauchi, A. and Miki, T. (2009). Intraspecific niche flexibility facilitates species coexistence in a competitive community with a fluctuating environment. Oikos, 118, 5566.Google Scholar
Yodzis, P. and Winemiller, K. O. (1999). In search of operational trophospecies in a tropical aquatic food web. Oikos, 87, 327340.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.

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.

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.

Available formats
×