Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-29T04:41:05.762Z Has data issue: false hasContentIssue false

Part III - Food Webs and Environmental Sustainability

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. 287 - 405
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

References

Albrecht, M., Riesen, M., and Schmid, B. (2010). Plant–pollinator network assembly along the chronosequence of a glacier foreland. Oikos, 119(10), 16101624.CrossRefGoogle Scholar
Alexander, J. M., Diez, J. M., and Levine, J. M. (2015). Novel competitors shape species’ response to climate change. Nature, 525, 515518.CrossRefGoogle Scholar
Amarasekare, P. (2008). Spatial dynamics of foodwebs. Annual Review of Ecology and Systematics, 39, 479500.Google Scholar
Araújo, M. B., Nogués-Bravo, D., Diniz-Filho, J. A. F., et al. (2008). Quaternary climate changes explain diversity among reptiles and amphibians. Ecography, 31, 815.CrossRefGoogle Scholar
Bascompte, J. (2009). Mutualistic networks. Frontiers in Ecology and the Environment, 8, 429436.CrossRefGoogle Scholar
Bascompte, J., Jordano, P., Melián, C. J., and Olesen, J. M. (2003). The nested assembly of plant–animal mutualistic networks. Proceedings of the National Academy of Sciences of the United States of America, 100, 93839387.Google Scholar
Bell, G. (2007). The evolution of trophic structure. Heredity, 99(5), 494505.CrossRefGoogle ScholarPubMed
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
Boulangeat, I., Gravel, D., and Thuiller, W. (2012). Accounting for dispersal and biotic interactions to disentangle the drivers of species distributions and their abundances. Ecology Letters, 15(6), 584593.Google Scholar
Briand, F. and Cohen, J. E. (1987). Environmental correlates of food chain length. Science, 238(4829), 956960.CrossRefGoogle ScholarPubMed
Cohen, J. E., Pimm, S. L., Yodzis, P., and Saldaña, J. (1993). Body sizes of animal predators and animal prey in food webs. Journal of Animal Ecology, 62, 6778.Google Scholar
Cornell, H. V. and Harrison, S. P. (2013). Regional effects as important determinants of local diversity in both marine and terrestrial systems. Oikos, 122, 288297.Google Scholar
Currie, D. J., Mittelbach, G. G., Cornell, H. V., et al. (2004). Predictions and tests of climate-based hypotheses of broad-scale variation in taxonomic richness. Ecology Letters, 7(12), 11211134.Google Scholar
Dalsgaard, B., Magård, E., Fjeldså, J., et al. (2011). Specialization in plant–hummingbird networks is associated with species richness, contemporary precipitation and quaternary climate-change velocity. PLoS ONE, 6(10), e25891.Google Scholar
Dalsgaard, B., Trøjelsgaard, K., Martín González, A. M., et al. (2013). Historical climate‐change influences modularity and nestedness of pollination networks. Ecography, 36(12), 13311340.Google Scholar
Darwin, C. R. (1862). On the Various Contrivances by Which British and Foreign Orchids Are Fertilised by Insects, and on the Good Effects of Intercrossing. London, UK: John Murray.Google Scholar
Dobzhansky, T. (1950). Evolution in the tropics. American Scientist, 38(2), 209221.Google Scholar
Dossena, M., Yvon-Durocher, G., Grey, J., et al. (2012). Warming alters community size structure and ecosystem functioning. Proceedings of the Royal Society B: Biological Sciences, 279, 30113019.Google Scholar
Dunne, J. A. (2006). The network structure of food webs. In Ecological Networks: Linking Structure to Dynamics in Food Webs, ed. Pascual, M. and Dunne, J. A., Oxford, UK: Oxford University Press, pp. 2786.Google Scholar
Dunne, J. A., Williams, R. J., and Martinez, N. D. (2002). Food-web structure and network theory: the role of connectance and size. Proceedings of the National Academy of Sciences of the United States of America, 99, 1291712922.Google Scholar
Gravel, D., Massol, F., Canard, E., Mouillot, D., and Mouquet, N. (2011). Trophic theory of island biogeography. Ecology Letters, 14(10), 10101016.Google Scholar
Guimaraes, P. R. Jr., Jordano, P., and Thompson, J. N. (2011). Evolution and coevolution in mutualistic networks. Ecology Letters, 14(9), 877885.Google Scholar
Hawkins, B. A. (2001). Ecology’s oldest pattern? Trends in Ecology and Evolution, 16, 470.Google Scholar
Hawkins, B. A., Porter, E. E., and Diniz-Filho, J. A. F. (2003). Productivity and history as predictors of the latitudinal diversity gradient of terrestrial birds. Ecology, 84, 16081623.CrossRefGoogle Scholar
Hawkins, B. A., Diniz‐Filho, J. A. F., Jaramillo, C. A., and Soeller, S. A. (2007). Climate, niche conservatism, and the global bird diversity gradient. American Naturalist, 170(S2), S16S27.Google Scholar
Holt, A. R., Warren, P. H., and Gaston, K. J. (2002). The importance of biotic interactions in abundance–occupancy relationships. Journal of Animal Ecology, 71(5), 846854.CrossRefGoogle Scholar
Holt, R. D. (1993). Ecology at the mesoscale: the influence of regional processes on local communities. In Species Diversity in Ecological Communities, ed. Ricklefs, R. and Schluter, D., Chicago, IL: University of Chicago Press, pp. 7788.Google Scholar
Holt, R. D. (1996). Food webs in space: an island biogeographic perspective. In Food Webs: Contemporary Perspectives, ed. Polis, G. A. and Winemiller, K., London: Chapman and Hall, pp. 313323.Google Scholar
Holt, R. D. (2010). Towards a trophic island biogeography: reflections on the interface of island biogeography and food web ecology. In The Theory of Island Biogeography Revisited, ed. Losos, J. B. and Ricklefs, R. E., Princeton, NJ: Princeton University Press, pp. 143185.Google Scholar
Holyoak, M., Leibold, M. A., and Holt, R. D. (2005). Metacommunities: Spatial Dynamics and Ecological Communities. Chicago, IL: University of Chicago Press.Google Scholar
Huston, M. A. (1999). Local processes and regional patterns: appropriate scales for understanding variation in the diversity of plants and animals. Oikos, 86, 393401.Google Scholar
Ings, T. C., Montoya, J. M., Bascompte, J., et al. (2009). Review: ecological networks – beyond food webs. Journal of Animal Ecology, 78(1), 253269.CrossRefGoogle ScholarPubMed
Jansson, R. and Dynesius, M. (2002). The fate of clades in a world of recurrent climatic change: Milankovitch oscillations and evolution. Annual Review of Ecology and Systematics, 33, 741777.Google Scholar
Janzen, D. H. (1973). Sweep samples of tropical foliage insects: effects of seasons, vegetation types, elevation, time of day, and insularity. Ecology, 54, 687708.Google Scholar
Johnson, D. H. (1980). The comparison of usage and availability measurements for evaluating resource preference. Ecology, 61(1), 6571.Google Scholar
Joppa, L. N., Bascompte, J., Montoya, J. M., et al. (2009). Reciprocal specialization in ecological networks. Ecology Letters, 12, 961969.Google Scholar
Jordano, P., Bascompte, J., and Olesen, J. M. (2003). Invariant properties in coevolutionary networks of plant–animal interactions. Ecology Letters, 6, 6981.Google Scholar
Kissling, W. D., Field, R., Korntheuer, H., Heyder, U., and Böhning-Gaese, K. (2010). Woody plants and the prediction of climate-change impacts on bird diversity. Philosophical Transactions of the Royal Society B: Biological Sciences, 365(1549), 20352045.CrossRefGoogle ScholarPubMed
Kissling, W. D., Dormann, C. F., Groeneveld, J., et al. (2012). Towards novel approaches to modelling biotic interactions in multispecies assemblages at large spatial extents. Journal of Biogeography, 39(12), 21632178.Google Scholar
Kitching, R. L. (2000). Food Webs and Container Habitats: the Natural History and Ecology of Phytotelmata. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Krause, A. E., Frank, K. A., Mason, D. M., Ulanowicz, R. E., and Taylor, W. W. (2003). Compartments revealed in food web structure, Nature, 426, 482485.CrossRefGoogle ScholarPubMed
Kreft, H. and Jetz, W. (2007). Global patterns and determinants of vascular plant diversity. Proceedings of the National Academy of Sciences of the United States of America, 104(14), 59255930.Google Scholar
Leibold, M. A., Holyoak, M., Mouquet, N., et al. (2004). The metacommunity concept: a framework for multi‐scale community ecology. Ecology Letters, 7(7), 601613.Google Scholar
Lurgi, M., Lopez, B. C., and Montoya, J. M. (2012a). Climate change impacts on body size and food web structure on mountain ecosystems. Philosophical Transactions of the Royal Society B: Biological Sciences, 367, 30503057.Google Scholar
Lurgi, M., Lopez, B. C., and Montoya, J. M. (2012b). Novel communities from climate change. Philosophical Transactions of the Royal Society B: Biological Sciences, 367, 29132922.Google Scholar
MacArthur, R. H. (1955). Fluctuations of animal populations, and a measure of community stability. Ecology, 36, 533536.Google Scholar
MacArthur, R. H. (1972). Geographical Ecology: Patterns in the Distribution of Species. New York, NY: Harper and Row.Google Scholar
MacArthur, R. H. and Pianka, E. R. (1966). On optimal use of a patchy environment. American Naturalist, 100, 603609.Google Scholar
MacArthur, R. H. and Wilson, E. O. (1967). The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
Martínez, N. D. (1991). Artifacts or attributes? Effects of resolution on the Little Rock Lake food web. Ecological Monographs, 64(4), 367392.Google Scholar
Massol, F., Gravel, D., Mouquet, N., et al. (2011). Linking community and ecosystem dynamics through spatial ecology. Ecology Letters, 14(3), 313323.Google Scholar
McCann, K. S., Rasmussen, J. B., and Umbanhowar, J. (2005). The dynamics of spatially coupled food webs. Ecology Letters, 8, 513523.CrossRefGoogle ScholarPubMed
McKane, A. J. and Drossel, B. (2005). Modelling evolving food webs. In Dynamic Food Webs: Multispecies Assemblages, Ecosystem Development, and Environmental Change, ed. de Ruiter, P. C., Wolters, V., and Moore, J. C., Oxford, UK: Academic Press, pp. 7488.Google Scholar
Melián, C. J. and Bascompte, J. (2004). Food web cohesion. Ecology, 85, 352358.CrossRefGoogle Scholar
Melián, C. J., Vilas, C., Baldó, F., González-Ortegón, E., Drake, P., and Williams, R. J. (2011). Eco-evolutionary dynamics of individual-based food webs. Advances in Ecological Research, 45, 225268.Google Scholar
Montoya, J. M. and Raffaelli, D. (2010). Climate change, biotic interactions and ecosystem services. Philosophical Transactions of the Royal Society B: Biological Sciences, 365(1549), 20132018.Google Scholar
Montoya, J. M., Rodriguez, M. Á., and Hawkins, B. A. (2003). Food web complexity and higher-level ecosystem services. Ecology Letters, 6, 587593.Google Scholar
Montoya, J. M., Pimm, S. L., and Solé, R V. (2006). Ecological networks and their fragility. Nature, 442, 259264.Google Scholar
Montoya, J. M., Woodward, G., Emmerson, M. C., and Solé, R. V. (2009). Press perturbations and indirect effects in real food webs. Ecology, 90, 24262433.Google Scholar
Morales-Castilla, I., Matias, M. G., Gravel, D., and Araújo, M. B. (2015). Inferring biotic interactions from proxies. Trends in Ecology and Evolution, 30, 347356.Google Scholar
Morris, R. J., Gripenberg, S., Lewis, O. T., and Roslin, T. (2014). Antagonistic interaction networks are structured independently of latitude and host guild. Ecology Letters, 17(3), 340349.Google Scholar
Moya-Laraño, J., Verdeny-Vilalta, O., Rowntree, J., et al. (2012). Chapter 1: Climate change and eco-evolutionary dynamics in food webs. Advances in Ecological Research, 47, 180.Google Scholar
Ollerton, J. and Cranmer, L. (2002). Latitudinal trends in plant‐pollinator interactions: are tropical plants more specialised? Oikos, 98(2), 340350.Google Scholar
Paine, R. T., Tegner, M. J., and Johnson, E. A. (1998). Compounded perturbations yield ecological surprises. Ecosystems, 1, 535545.Google Scholar
Pearson, R. G. and Dawson, T. P. (2003). Predicting the impacts of climate change on the distribution of species: are bioclimate envelope models useful? Global Ecology and Biogeography, 12(5), 361371.Google Scholar
Polis, G. A. (1991). Complex trophic interactions in deserts: an empirical critique of food web theory. American Naturalist, 138, 123155.Google Scholar
Post, D. M. (2002). The long and short of food-chain length. Trends in Ecology and Evolution, 17(6), 269277.Google Scholar
Post, D. M., Pace, M. L., and Hairston, N. G. Jr (2000). Ecosystem size determines food-chain length in lakes. Nature, 405, 10471049.Google Scholar
Rall, B. C., Vucic-Pestic, O., Ehnes, R. B., Emmerson, M., and Brose, U. (2010). Temperature, predator–prey interaction strength and population stability. Global Change Biology, 16(8), 21452157.CrossRefGoogle Scholar
Reiss, J., Bridle, J. R., Montoya, J. M., and Woodward, G. (2009). Emerging horizons in biodiversity and ecosystem functioning research. Trends in Ecology and Evolution, 24(9), 505514.Google Scholar
Ricklefs, R. E. (1987). Community diversity: relative roles of local and regional processes. Science, 235, 167171.Google Scholar
Romdal, T. S., Araújo, M. B., and Rahbek, C. (2013). Life on a tropical planet: niche conservatism and the global diversity gradient. Global Ecology and Biogeography, 22(3), 344350.Google Scholar
Rooney, N., McCann, K. S., Gellner, G., and Moore, J. (2006). Structural asymmetry and the stability of diverse food webs. Nature, 442, 265269.Google Scholar
Rooney, N., McCann, K. S., and Moore, J. C. (2008). A landscape theory for food web architecture. Ecology Letters, 11(8), 867881.Google Scholar
Sagarin, R. D., Gaines, S. D., and Gaylord, B. (2006). Moving beyond assumptions to understand abundance distributions across the ranges of species. Trends in Ecology and Evolution, 21(9), 524530.Google Scholar
Schemske, D. W. (2002). Ecological and evolutionary perspectives on the origins of tropical diversity. In Foundations of Tropical Forest Biology, ed. Chazdon, R. L. and Whitmore, T. C., Chicago, IL: University of Chicago Press, pp. 163173.Google Scholar
Schemske, D. W., Mittelbach, G. G., Cornell, H. V., Sobel, J. M., and Roy, K. (2009). Is there a latitudinal gradient in the importance of biotic interactions? Annual Review of Ecology and Systematics, 40, 245269.Google Scholar
Schleuning, M., Fründ, J., Klein, A.-M., et al. (2012). Specialization of mutualistic interaction networks decreases toward tropical latitudes. Current Biology, 22, 19251931.Google Scholar
Shurin, J. B., Clasen, J. L., Greig, H. S., Kratina, P., and Thompson, P. L. (2012). Warming shifts top–down and bottom–up control of pond food web structure and function. Philosophical Transactions of the Royal Society B: Biological Sciences, 367, 30083017.Google Scholar
Solé, R. V. and Montoya, J. M. (2001). Complexity and fragility in ecological networks. Proceedings of the Royal Society B: Biological Sciences, 268, 20392045.Google Scholar
Stevens, G. C. (1989). The latitudinal gradients in geographical range: how so many species co-exist in the tropics. American Naturalist, 133, 240256.Google Scholar
Thompson, R. M., Brose, U., Dunne, J. A., et al. (2012). Food webs: reconciling the structure and function of biodiversity. Trends in Ecology and Evolution, 27(12), 689697.Google Scholar
Thuiller, W., Brotons, L., Araújo, M. B., and Lavorel, S. (2004). Effects of restricting environmental range of data to project current and future species distributions. Ecography, 27(2), 165172.Google Scholar
Trøjelsgaard, K. and Olesen, J. M. (2013). Macroecology of pollination networks. Global Ecology and Biogeography, 22(2), 149162.Google Scholar
Tylianakis, J. M., Tscharntke, T., and Lewis, O. T. (2007). Habitat modification alters the structure of tropical host–parasitoid food webs. Nature, 445(7124), 202205.Google Scholar
Tylianakis, J. M., Didham, R. K., Bascompte, J., and Wardle, D. A. (2008). Global change and species interactions in terrestrial ecosystems. Ecology Letters, 11(12), 13511363.Google Scholar
Vasseur, D. A. and McCann, K. S. (2005). A mechanistic approach for modeling temperature‐dependent consumer–resource dynamics. American Naturalist, 166(2), 184198.Google Scholar
Vázquez, D. P. and Stevens, R. D. (2004). The latitudinal gradient in niche breadth: concepts and evidence. American Naturalist, 164(1), E119.Google Scholar
Warren, P. H. (1990). Variation in food-web structure: the determinants of connectance. American Naturalist, 136(5), 689700.Google Scholar
Wisz, M. S., Pottier, J., Kissling, W. D., et al. (2013). The role of biotic interactions in shaping distributions and realised assemblages of species: implications for species distribution modelling. Biological Reviews, 88(1), 1530.Google Scholar
Wootton, J. T. and Emmerson, M. (2005). Measurement of interaction strength in nature. Annual Review of Ecology, Evolution, and Systematics, 5, 419444.Google Scholar
Yvon-Durocher, G., Jones, J., Trimmer, M., Woodward, G., and Montoya, J. M. (2010). Warming alters the metabolic balance of ecosystems. Philosophical Transactions of the Royal Society B: Biological Sciences, 365(1549), 21172126.Google Scholar
Yvon-Durocher, G., Montoya, J. M., Woodward, G., Jones, J., and Trimmer, M. (2011). Warming increases the proportion of primary production emitted as methane from freshwater mesocosms. Global Change Biology, 17(2), 12251234.Google Scholar

References

Alcántara, J. M. and Rey, P. J. (2012). Linking topological structure and dynamics in ecological networks. American Naturalist, 180, 186199.Google Scholar
Allain, V. (2005). Ecopath model of the pelagic ecosystem of the Western and Central Pacific Ocean. First Regular Session of the Scientific Committee of the Western and Central Pacific Fisheries Commission, WCPFC-SC1-EB WP-10, 119.Google Scholar
Allain, V., Nicol, S., Essington, T., et al. (2007). An Ecopath with Ecosim model of the Western and Central Pacific Ocean warm pool pelagic ecosystem. Third Regular Session of the Scientific Committee of the Western and Central Pacific Fisheries Commission, WCPFC-SC3-EB SWG/IP-8, 142.Google Scholar
Allesina, S. and Pascual, M. (2008). Network structure, predator–prey modules, and stability in large food webs. Theoretical Ecology, 1, 5564.Google Scholar
Andrewartha, H. G. and Birch, C. (1954). The Distribution and Abundance of Animals. Chicago, IL: University of Chicago Press.Google Scholar
Angert, A. L., Huxman, T. E., Chesson, P., and Venable, D. L. (2009). Functional tradeoffs determine species coexistence via the storage effect. Proceedings of the National Academy of Sciences of the United States of America, 106, 1164111645.Google Scholar
Atkinson, A., Hill, S. L., Barange, M., et al. (2014). Sardine cycles, krill declines, and locust plagues: revisiting “wasp-waist” food webs. Trends in Ecology and Evolution, 29, 309316.Google Scholar
Bailly, D., Agostinho, A. A., and Suzuki, H. I. (2008). Influence of the flood regime on the reproduction of fish species with different reproductive strategies in the Cuiabá River, Upper Pantanal, Brazil. River Research and Applications, 24, 12181229.Google Scholar
Bakun, A. and Broad, K. (2003). Environmental “loopholes” and fish population dynamics: comparative pattern recognition with focus on El Niño effects in the Pacific. Fisheries Oceanography, 12, 458473.Google Scholar
Barton, B. T. and Ives, A. R. (2014). Species interactions and a chain of indirect effects driven by reduced precipitation. Ecology, 95, 486494.Google Scholar
Bellmore, J. R., Baxter, C. V., and Connolly, P. J. (2015). Spatial complexity reduces interaction strengths in the meta-food web of a river floodplain mosaic. Ecology, 96, 274283.Google Scholar
Camerano, L. (1880). Dell’equilibrio dei viventi merce la reciproca distribuzione. Atti della Scienze di Torino, 15, 393414.Google Scholar
Carpenter, S. R. and Turner, M. G. (2001). Hares and tortoises: interactions of fast and slow variables in ecosystems. Ecosystems, 3, 495497.Google Scholar
Cheal, A. J., Delean, S., and Thompson, A. A. (2007). Spatial synchrony in coral reef fish populations and the influence of climate. Ecology, 88, 158169.Google Scholar
Chesson, P. (2000). General theory of competitive coexistence in spatially varying environments. Theoretical Population Biology, 58, 211237.Google Scholar
Chesson, P. and Warner, R. (1981). Environmental variability promotes coexistence in lottery competitive systems. American Naturalist, 117, 923943.Google Scholar
Christensen, V. (2013). Ecological networks in fisheries: predicting the future? Fisheries, 38, 7681.Google Scholar
Christensen, V. and Walters, C. J. (2004). Ecopath with Ecosim: methods, capabilities and limitations. Ecological Modelling, 172(2–4), 109139.Google Scholar
De Roos, A. M., Persson, L., and McCauley, E. (2003). The influence of size-dependent life history traits on the structure and dynamics of populations and communities. Ecology Letters, 6, 473487.Google Scholar
Fujiwara, M. (2016). Incorporating demographic diversity into food web models: effects on community structure and dynamics. Ecological Modelling, 322, 1018.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
Grime, J. P. (1977). Evidence for existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. American Naturalist, 111, 11691194.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
Hartvig, M., Andersen, K. H., and Beyer, J. E. (2011). Food web framework for size-structured populations. Journal of Theoretical Biology, 272, 113122.Google Scholar
Hastings, A. (1996). What equilibrium behavior of Lotka–Volterra models does not tell us about food webs. In Food Webs: Integration of Patterns and Dynamics, ed. Polis, G. A. and Winemiller, K. O., New York, NY: Chapman & Hall, pp. 211217.Google Scholar
Hastings, A., Hom, C., Ellner, S., Turchin, P., and Godfray, H. C. J. (1993). Chaos in ecology: is mother nature a strange attractor? Annual Reviews of Ecology and Systematics, 24, 133.Google Scholar
Haydon, D. (1994). Pivotal assumptions determining the relationship between stability and complexity: an analytical synthesis of the stability-complexity debate. American Naturalist, 144, 1429.Google Scholar
Higgins, S., Pickett, S. T. A., and Bond, W. J. (2000). Predicting extinction risks for plants: environmental stochasticity can save declining populations. Trends in Ecology and Evolution, 15, 516520.Google Scholar
Holt, R. D. (2008). Theoretical perspectives on resource pulses. Ecology, 89, 671681.Google Scholar
Holt, R. D. (2009). Towards a trophic island biogeography: reflections on the interface of island biogeography and food web ecology. In The Theory of Island Biogeography Revisited, ed. Losos, J. B. and Ricklefs, R. E., Princeton, NJ: Princeton University Press, pp. 143185.CrossRefGoogle Scholar
Jacobsen, N. S., Gislason, H., and Andersen, K. H. (2014). The consequences of balanced harvesting of fish communities. Proceedings of the Royal Society B: Biological Sciences, 281(1775), DOI: 10.1098/rspb.2013.2701.Google Scholar
Koenig, W. D. and Liebhold, A. M. (2005). Effects of periodical cicada emergences on abundance and synchrony of avian populations. Ecology, 86, 18731882.Google Scholar
Lawton, J. H. (1995). Population dynamic principles. In Extinction Rates, ed. Lawton, J. H. and May, R. M., Oxford, UK: Oxford University Press, pp. 147163.Google Scholar
Lindeman, R. L. (1942). The trophic-dynamic aspect of ecology. Ecology, 23, 399418.Google Scholar
Madsen, T. and Shine, R. (2000). Rain, fish and snakes: climatically driven population dynamics of Arafura filesnakes in tropical Australia. Oecologia, 124, 208215.Google Scholar
May, R. M. (1973). Stability and Complexity in Model Ecosystems. Princeton, NJ: Princeton University Press.Google Scholar
McCann, K. (2012). Food Webs. Princeton, NJ: Princeton University Press.Google Scholar
Melián, C. J., Baldó, F., Matthews, B., et al. (2014). Individual trait variation and diversity in food webs. Advances in Ecological Research, 50, 207241.Google Scholar
Menge, B. A., Gouhier, T., Friedenburg, T., et al. (2011). Linking long-term, large-scale climatic and environmental variability to patterns of marine invertebrate recruitment: toward explaining “unexplained” variation. Journal of Experimental Marine Biology and Ecology, 400, 236249.CrossRefGoogle Scholar
Olff, H., Alonso, D., Berg, M. P., et al. (2009). Parallel ecological networks in ecosystems. Philosophical Transactions of the Royal Society B: Biological Sciences, 364, 17551779.Google Scholar
Paine, R. T. (1988). On food webs: road maps of interactions or the grist for theoretical development? Ecology, 69, 16481654.Google Scholar
Pianka, E. R. (1970). On r- and K-selection. American Naturalist, 104, 592597.Google Scholar
Polis, G. A., Holt, R. D., Menge, B. A., and Winemiller, K. O. (1996). Time, space, and life history: influences on food webs. In Food Webs: Integration of Patterns and Dynamics, ed. Polis, G. A. and Winemiller, K. O., New York, NY: Chapman and Hall, pp. 435460.Google Scholar
Post, D. M. (2002). The long and short of food-chain length. Trends in Ecology and Evolution, 17, 269277.CrossRefGoogle Scholar
Power, M. E., Parker, M. S., and Dietrich, W. E. (2008). Seasonal reassembly of a river food web: floods, droughts, and impacts of fish. Ecological Monographs, 78, 263282.Google Scholar
Rose, K. A. (2012). End-to-end models for marine ecosystems: are we on the precipice of a significant advance or just putting lipstick on a pig? Scientia Marina, 76, 195201.Google Scholar
Rose, K. A., Cowan, J. H., Winemiller, K. O., Myers, R. A., and Hilborn, R. (2001). Compensatory density-dependence in fish populations: importance, controversy, understanding, and prognosis. Fish and Fisheries, 2, 293327.Google Scholar
Rudolf, V. H. W. and Lafferty, K. D. (2011). Stage structure alters how complexity affects stability of ecological networks. Ecology Letters, 14, 7579.Google Scholar
Schmitz, O. J. and Booth, G. (1997). Modelling food web complexity: the consequences of individual-based, spatially explicit behavioural ecology on trophic interactions. Evolutionary Ecology, 11, 379398.Google Scholar
Siddon, E. C., Kristiansen, T., Mueter, F. J., et al. (2013). Spatial match-mismatch between juvenile fish and prey provides a mechanism for recruitment variability across contrasting climate conditions in the Eastern Bering Sea. PLOS One, 8(12), e84526.Google Scholar
Sinclair, A. R. E. (2003). Mammal population regulation, keystone processes and ecosystem dynamics. Philosophical Transactions of the Royal Society B: Biological Sciences, 358, 17291740.Google Scholar
Soberón, J. (2007). Grinnellian and Eltonian niches and geographic distributions of species. Ecology Letters, 10, 11151123.Google Scholar
Southwood, T. R. E. (1977). Habitat, the templet for ecological strategies? Journal of Animal Ecology, 46, 337365.Google Scholar
Stearns, S. C. (1992). The Evolution of Life Histories. Oxford, UK: Oxford University Press.Google Scholar
Summerhayes, V. S. and Elton, C. S. (1923). Contribution to the ecology of Spitsbergen and Bear Island. Journal of Ecology, 11, 214286.Google Scholar
Tang, S., Pawar, S., and Allesina, S. (2014). Correlation between interaction strengths drives stability in large ecological networks. Ecology Letters, 17, 10941100.Google Scholar
Travers, M., Shin, Y.-J., Jennings, S., and Cury, P. (2007). Towards end-to-end models for investigating the effects of climate and fishing in marine ecosystems. Progressive Oceanography, 75, 751770.Google Scholar
Travis, J., Coleman, F. C., Auster, P. J., et al. (2014). Integrating the invisible fabric of nature into fisheries management. Proceedings of the National Academy of Sciences of the United States of America, 111(2), 581584.Google Scholar
Turchin, P. (2003). Complex Population Dynamics. A Theoretical/Empirical Synthesis. Princeton, NJ: Princeton University Press.Google Scholar
van Kooten, T., de Roos, A. M., and Persson, L. (2005). Bistability and an Allee effect as emergent consequences of stage-specific predation. Journal of Theoretical Biology, 237, 6774.Google Scholar
Van Winkle, W., Rose, K. A., and Chambers, R. C. (1993). Individual-based approach to fish population dynamics: an overview. Transactions of the American Fisheries Society, 122, 397403.2.3.CO;2>CrossRefGoogle Scholar
Varughese, M. M. (2011). A framework for modelling ecological communities and their interactions with the environment. Ecological Complexity, 8, 105112.Google Scholar
Vasseur, D. and Gaedke, U. (2007). Spectral analysis unmasks synchronous and compensatory dynamics in plankton communities. Ecology, 88, 20582071.Google Scholar
Winemiller, K. O. (1989). Patterns of variation in life history among South American fishes in seasonal environments. Oecologia, 81, 225241.Google Scholar
Winemiller, K. O. (1990). Spatial and temporal variation in tropical fish trophic networks. Ecological Monographs, 60, 331367.Google Scholar
Winemiller, K. O. (1992). Life history strategies and the effectiveness of sexual selection. Oikos, 62, 318327.Google Scholar
Winemiller, K. O. (2005). Life history strategies, population regulation, and their implications for fisheries management. Canadian Journal of Fisheries and Aquatic Sciences, 62, 872885.CrossRefGoogle Scholar
Winemiller, K. O. and Rose, K. A. (1992). Patterns of life-history diversification in North American fishes: implications for population regulation. Canadian Journal of Fisheries and Aquatic Sciences, 49, 21962218.Google Scholar
Winemiller, K. O., Roelke, D. L., Cotner, J. B., et al. (2014). Top–down control of basal resources in a cyclically pulsing ecosystem. Ecological Monographs, 84, 621635.Google Scholar
Woodson, C. B., McManus, M. A., Tyburczy, J. A., et al. (2012). Coastal fronts set recruitment and connectivity patterns across multiple taxa. Limnology and Oceanography, 57, 582596.Google Scholar
Worm, B. and Duffy, J. E. (2003). Biodiversity, productivity and stability in real food webs. Trends in Ecology and Evolution, 18, 628632.Google Scholar
Yang, L. H. and Rudolf, V. H. W. (2010). Phenology, ontogeny and the effects of climate change on the timing of species interactions. Ecology Letters, 13, 110.Google Scholar
Yodzis, P. and Innes, S. (1992). Body size and consumer-resource dynamics. American Naturalist, 139, 11511175.Google Scholar
Zeglin, L. H., Bottomley, P. J., Jumpponen, A., et al. (2013). Altered precipitation regime affects the function and composition of soil microbial communities on multiple time scales. Ecology, 94, 23342345.Google Scholar
Zhou, C., Fujiwara, M., and Grant, W. E. (2013). Dynamics of a predator–prey interaction with seasonal reproduction and continuous predation. Ecological Modelling, 268, 2536.Google Scholar

References

Allesina, S. and Tang, S. (2012). Stability criteria for complex ecosystems. Nature, 483, 205208.Google Scholar
Barnosky, A. D., Matzke, N., Tomiya, S., et al. (2011). Has the Earth’s sixth mass extinction already arrived? Nature, 471, 5157.CrossRefGoogle ScholarPubMed
Beddington, J. R., Free, C. A., and Lawton, J. H. (1978). Characteristics of successful natural enemies in models of biological control of insect pests. Nature, 273, 513519.CrossRefGoogle ScholarPubMed
Bender, E. A., Case, T. J., and Gilpin, M. E. (1984). Perturbation experiments in community ecology: theory and practice. Ecology, 65, 113.Google Scholar
Berg, S., Christianou, M., Jonsson, T., and Ebenman, B. (2011). Using sensitivity analysis to identify keystone species and keystone links in size-based food webs. Oikos, 120, 510519.CrossRefGoogle Scholar
Berg, S., Pimenov, A., Palmer, C., Emmerson, M. C., and Jonsson, T. (2015). Ecological communities are vulnerable to realistic extinction sequences. Oikos, 124, 486496.Google Scholar
Berlow, E. L., Dunne, J. A., Martinez, N. D, et al. (2009). Simple prediction of interaction strengths in compex food webs. PNAS, 6, 187191.Google Scholar
Borrvall, C. and Ebenman, B. (2006). Early onset of secondary extinctions in ecological communities following the loss of top predators. Ecology Letters, 9, 435442.Google Scholar
Borrvall, C., Ebenman, B., and Jonsson, T. (2000). Biodiversity lessens the risk of cascading extinction in model food webs. Ecology Letters, 3, 131136.Google Scholar
Bracken, M. E. S. and Low, N. H. N. (2012). Realistic losses of rare species disproportionately impact higher trophic levels. Ecology Letters, 15, 461467.Google Scholar
Brown, J. H., Gillooly, J. F., Allen, A. P., and Savage, V. M. (2004). Toward a metabolic theory of ecology. Ecology, 85, 17711789.Google Scholar
Case, T. J. (2000). An Illustrated Guide to Theoretical Ecology. Oxford, UK: Oxford University Press.Google Scholar
Chesson, P. and Kuang, J. J. (2008). The interaction between predation and competition. Nature, 456, 235238.Google Scholar
Cottee-Jones, H. E. W. and Whittaker, R. J. (2012). The keystone species concept: a critical appraisal. Frontiers of Biogeography, 4, 117127.Google Scholar
Curtsdotter, A., Binzer, A., Brose, U., et al. (2011). Robustness to secondary extinctions: comparing trait-based sequential deletions in static and dynamic food webs. Basic and Applied Ecology, 12, 571580.Google Scholar
Ebenman, B. and Jonsson, T. (2005). Using community viability analysis to identify fragile systems and keystone species. Trends in Ecology and Evolution, 20, 568575.Google Scholar
Eklöf, A. and Ebenman, B. (2006). Species loss and secondary extinctions in simple and complex model communities. Journal of Animal Ecology, 75, 239246.Google Scholar
Ellison, A. M., Bank, M. S., Clinton, B. D., et al. (2005). Loss of foundation species: consequences for the structure and dynamics of forested ecosystems. Frontiers in Ecology and the Environment, 3, 479486.Google Scholar
Estes, J. A., Terborgh, J., Brashares, J. S., et al. (2011). Trophic downgrading of planet Earth. Science, 333, 301306.Google Scholar
Fowler, M. (2010). Extinction cascades and the distribution of species interactions. Oikos, 119, 864873.Google Scholar
Gaston, K. J. (2010). Valuing common species. Science, 327, 154155.CrossRefGoogle ScholarPubMed
Gaston, K. J. (2011). Common ecology. BioScience, 61, 354362.Google Scholar
Gaston, K. J. and Fuller, R. A. (2008). Commonness, population depletion and conservation biology. Trends in Ecology and Evolution, 23, 1419.Google Scholar
Hol, W. H. G., de Boer, W., Termorshuizen, A. J., et al. (2010). Reduction of rare soil microbes modifies plant-herbivore interactions. Ecology Letters, 13, 292301.Google Scholar
Hooper, D. U., Adair, E. C., Cardinale, B. J., et al. (2012). A global synthesis reveals biodiversity loss as a major driver of ecosystem change. Nature, 486, 105108.Google Scholar
Hubbell, S. P. (2001). The Unified Neutral Theory of Biodiversity and Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
Isbell, F., Calcagno, V., Hector, A., et al. (2011). High plant diversity is needed to maintain ecosystem services. Nature, 477, 199202.CrossRefGoogle ScholarPubMed
Ives, A. R. and Cardinale, B. J. (2004). Food-web interactions govern the resistance of communities after non-random extinctions. Nature, 429, 174177.Google Scholar
Jonsson, T., Cohen, J. E., and Carpenter, S. R. (2005). Food webs, body size, and species abundance in ecological community description. Advances in Ecological Research, 36, 184.Google Scholar
Jonsson, T., Berg, S., Pimenov, A., and Emmerson, M. C. (2015). The context dependency of species keystone status during food web disassembly. Food Webs, 5, 110.Google Scholar
Lande, R. (1993). Risks of population extinction from demographic and environmental stochasticity and random catastrophes. American Naturalist, 142, 911927.Google Scholar
Lyons, K. G. and Schwartz, M. W. (2001). Rare species loss alters ecosystem function: invasion resistance. Ecology Letters, 4, 358365.Google Scholar
Lyons, K. G., Brigham, C. A., Traut, B. H., and Schwartz, M. W. (2005). Rare species and ecosystem functioning. Conservation Biology, 19, 10191024.Google Scholar
MacArthur, R. H. and Wilson, E. O. (1967). The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
May, R. M. (1975). Patterns of species abundance and diversity. In Ecology and Evolution of Communities, ed. Cody, M. L. and Diamond, J. M., Cambridge, UK: Belknap Press, pp. 81120.Google Scholar
McCoy, M. W. and Gillooly, J. F. (2008). Predicting natural mortality rates of plants and animals. Ecology Letters, 11, 710716.Google Scholar
Montoya, J. M., Emmerson, M. C., Solé, R. V., and Woodward, G. (2005). Perturbations and indirect effects in complex food webs. In Dynamic Food Webs: Multispecies Assemblages, Ecosystem Development and Environmental Change, ed. de Ruiter, P. C., Wolters, V., and Moore, J. C., Burlington, MA: Academic Press, pp. 369380.Google Scholar
Montoya, J. M., Pimm, S. L., and Sole, R. V. (2006). Ecological networks and their fragility. Nature, 442, 259264.Google Scholar
Montoya, J. M., Woodward, G., Emmerson, M. C., and Solé, R. V. (2009). Press perturbations and indirect effects in real food webs. Ecology, 90, 24262433.Google Scholar
Morin, P. (2011). Community Ecology. Oxford, UK: Wiley-Blackwell.Google Scholar
Mougi, A. and Kondoh, M. (2012). Diversity of interaction types and ecological community stability. Science, 337, 349351.Google Scholar
Novak, M. (2010). Estimating interaction strengths in nature: experimental support for an observational approach. Ecology, 91, 23942405.Google Scholar
O’Gorman, E. J., Jacob, U., Jonsson, T., and Emmerson, M. C. (2010). Interaction strength, food web topology and the relative importance of species in food webs. Journal of Animal Ecology, 79, 682692.Google Scholar
Paine, R. T. (1966). Food web complexity and species diversity. American Naturalist, 100, 6575.Google Scholar
Paine, R. T. (1969). A note on trophic complexity and community stability. American Naturalist, 355, 7375.Google Scholar
Pereira, H. M., Leadley, P. W., Proenca, V., et al. (2010). Scenarios for global biodiversity in the 21st century. Science, 330, 14961501.Google Scholar
Petchey, O., Eklöf, A., Borrvall, C., and Ebenman, B. (2008). Trophically unique species are vulnerable to cascading extinction. American Naturalist, 171, 568579.Google Scholar
Peters, R. H. (1983). The Ecological Implications of Body Size. New York, NY: Cambridge University Press.Google Scholar
Pinheiro, J. C. and Bates, D. M. (2004). Mixed-Effects Models in S and S-PLUS. New York, NY: Springer-Verlag.Google Scholar
Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D., et al. (2012). nlme: Linear and Nonlinear Mixed Effects Models. R package version 3.1–104.Google Scholar
R Development Core Team (2012). R: A Language and Environment for Statistical Computing. Vienna, Austria: R Foundation for Statistical Computing.Google Scholar
Reuman, D. C., Mulder, C., Raffaelli, D., and Cohen, J. E. (2008). Three allometric relations of population density to body mass: theoretical integration and empirical tests in 149 food webs. Ecology Letters, 11, 12161228.Google Scholar
Saavedra, S., Stouffer, D. B., Uzzi, B., and Bascompte, J. (2011). Strong contributers to network persistence are the most vulnerable to extinction. Nature, 478, 233236.Google Scholar
Sahasrabudhe, S. and Motter, A. E. (2011). Rescuing ecosystems from extinction cascades through compensatory perturbations. Nature Communications, 2, 18.Google Scholar
Säterberg, T., Sellman, S., and Ebenman, B. (2013). High frequency of functional extinctions in ecological networks. Nature, 499, 468470.Google Scholar
Thompson, K. (2010). Do We Need Pandas? Foxhole, Dartington: Green Books.Google Scholar
Wood, S. A., Lilley, S. A., Schiel, D. R., and Shurin, J. B. (2010). Organismal traits are more important than environment for species interactions in the intertidal zone. Ecology Letters, 13, 11601171.Google Scholar
Wootton, T. and Emmerson, M. (2005). Measurement of interaction strength in nature. Annual Review of Ecology, Evolution and Systematics, 36, 419444.Google Scholar
Yodzis, P. and Innes, S. (1992). Body size and consumer-resource dynamics. American Naturalist, 139, 11511175.Google Scholar

References

Berlow, E. L. (1999). Strong effects of weak interactions in ecological communities. Nature, 398, 330334.Google Scholar
Bolnick, D. I., Amarasekare, P., Araújo, M. S., et al. (2011). Why intraspecific trait variation matters in community ecology. Trends in Ecology and Evolution, 26, 183192.Google Scholar
Bondavalli, C., Bodini, A., Rossetti, G., and Allesina, S. (2006). Detecting stress at the whole-ecosystem level: the case of a mountain lake (Lake Santo, Italy). Ecosystems, 9, 768787.CrossRefGoogle Scholar
Dematté, L., Priami, C., Romanel, A., et al. (2007). BetaWB: modelling and simulating biological processes. In Proceedings of Summer Computer Simulation Conference (SCSC 2007), ed. Wainer, G. A. and Vakilzadian, H., pp. 777784.Google Scholar
Dematté, L., Priami, C., Romanel, A., et al. (2008). The Beta Workbench: a computational tool to study the dynamics of biological systems. Briefings in Bioinformatics, 9, 437449.Google Scholar
de Ruiter, P. C., Wolters, V., and Moore, J. C. (2005). Dynamic Food Webs: Multispecies Assemblages, Ecosystem Development and Environmental Change. Amsterdam: Academic Press.Google Scholar
Dunne, J. A. (2006). The network structure of food webs. In Ecological Networks: Linking Structure to Dynamics in Food Webs, ed. Pascual, M. and Dunne, J. A., Oxford: Oxford University Press, pp. 2786.Google Scholar
Eggers, S. L., Lewandowska, A. M., Barcelos e Ramos, J., et al. (2014). Community composition has greater impact on the functioning of marine phytoplankton communities than ocean acidification. Global Change Biology, 20, 713723.Google Scholar
Feest, A., Aldred, T. D., and Jedamzik, K. (2010). Biodiversity quality: a paradigm for biodiversity. Ecological Indicators, 10, 10771082.Google Scholar
Forlin, M. (2010). Knowledge discovery for stochastic models of biological systems. University of Trento, Ph.D. Thesis.Google Scholar
Gillespie, D. T. (1977). Exact stochastic simulation of coupled chemical reactions. Journal of Physical Chemistry, 81, 23402361.Google Scholar
Grimm, V. and Railsback, S. F. (2005). Individual-Based Modeling and Ecology. Princeton, NJ: Princeton University Press.Google Scholar
Hurlbert, S. H. (1997). Functional importance vs keystoneness: reformulating some questions in theoretical biocenology. Australian Journal of Ecology, 22, 369382.Google Scholar
Jordán, F., Scotti, M., and Priami, C. (2011). Process algebra-based models in systems ecology. Ecological Complexity, 8, 357363.Google Scholar
Jordán, F., Gjata, N., Mei, S., and Yule, C. M. (2012). Simulating food web dynamics along a gradient: quantifying human influence. PLoS ONE, 7(7), e40280.Google Scholar
Judson, O. P. (1994). The rise of the individual-based model in ecology. Trends in Ecology & Evolution, 9, 914.Google Scholar
Kennedy, J. and Eberhart, R. (1995). Particle swarm optimization. Proceedings of IEEE International Conference on Neural Networks IV, pp. 19421948.Google Scholar
Lafferty, K. D. and Dunne, J. A. (2010). Stochastic ecological network occupancy (SENO) models: a new tool for modelling ecological networks across spatial scales. Theoretical Ecology, 3, 123135.Google Scholar
Lande, R. (1988). Genetics and demography in biological conservation. Science, 241, 14551460.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
Montoya, J. M. and Solé, R. V. (2002). Small world patterns in food webs. Journal of Theoretical Biology, 214, 405412.Google Scholar
Priami, C. (2009). Algorithmic systems biology. Communications of ACM, 52, 8089.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.Google Scholar
Siokou-Frangou, I., Bianchi, M., Christaki, U., et al. (2002). Carbon flow in the planktonic food web along a gradient of oligotrophy in the Aegean Sea (Mediterranean Sea). Journal of Marine Systems, 33–34, 335353.Google Scholar
Tallberg, P., Horppila, J., Väisänen, A., and Nurminen, L. (1999). Seasonal succession of phytoplankton and zooplankton along a trophic gradient in a eutrophic lake: implications for food web management. Hydrobiologia, 412, 8194.Google Scholar
Ulanowicz, R. E. (1996). Trophic flow networks as indicators of ecosystem stress. In Food Webs: Integration of Patterns and Dynamics, ed. Polis, G. A. and Winemiller, K. O., London: Chapman and Hall, pp. 358368.Google Scholar
Valentini, R. and Jordán, F. (2010). CoSBiLab Graph: the network analysis module of CoSBiLab. Environmental Modeling Software, 25, 886888.Google Scholar
Warren, P. H. (1989). Spatial and temporal variation in the structure of a freshwater food web. Oikos, 55, 299311.Google Scholar
Winemiller, K. O. (1996). Factors driving temporal and spatial variation in aquatic floodplain food webs. In Food Webs: Integration of Patterns and Dynamics, ed. Polis, G. A. and Winemiller, K. O., London: Chapman and Hall, pp. 298312.Google Scholar
Yule, C. M. (1995). The impact of sediment pollution on the benthic invertebrate fauna of the Kelian River, East Kalimantan, Indonesia. In Tropical Limnology, Vol. III, ed. Timotius, K. H. and Göltenboth, F., Salatiga, Indonesia: SatyaWacana University Press, pp. 6175.Google Scholar
Yule, C. M., Boyero, L., and Marchant, R. (2010). Effects of sediment pollution on food webs in a tropical river (Borneo, Indonesia). Marine and Freshwater Research, 61, 204213.Google Scholar

References

Alfaro, A. C., Thomas, F., Sergent, L., and Duxbury, M. (2006). Identification of trophic interactions within an estuarine food web (northern New Zealand) using fatty acid biomarkers and stable isotopes. Estuarine, Coastal and Shelf Science, 70(1), 271286.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(5), 321329.Google Scholar
Bascompte, J., Melián, C. J., and Sala, E. (2005). Interaction strength combinations and the overfishing of a marine food web. Proceedings of the National Academy of Sciences of the United States of America, 102(15), 54435447.Google Scholar
Bhadra, A., Jordán, F., Sumana, A., Deshpande, S. A., and Gadagkar, R. (2009). A comparative social network analysis of wasp colonies and classrooms: linking network structure to functioning. Ecological Complexity, 6(1), 4855.Google Scholar
Black, A. J. and McKane, A. J. (2012). Stochastic formulation of ecological models and their applications. Trends in Ecology and Evolution, 27(6), 337345.Google Scholar
Bolnick, D. I., Svanbäck, R., Fordyce, J. A., et al. (2003). The ecology of individuals: incidence and implications of individual specialization. American Naturalist, 161(1), 128.Google Scholar
Bolnick, D. I., Amarasekare, P., Araújo, M. S., et al. (2011). Why intraspecific trait variation matters in community ecology. Trends in Ecology and Evolution, 26(4), 183192.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(3), 849872.Google Scholar
Brown, J. S. and Pavlovic, N. B. (1992). Evolution in heterogeneous environments: effects of migration on habitat specialization. Evolutionary Ecology, 6(5), 360382.Google Scholar
Croft, D. P., James, R., Ward, A. J. W., et al. (2005). Assortative interactions and social networks in fish. Oecologia, 143(2), 211219.Google Scholar
Crooks, K. R. and Soulé, M. E. (1999). Mesopredator release and avifaunal extinctions in a fragmented system. Nature, 400(6744), 563566.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.Google 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(1), 113.Google Scholar
Dematté, L., Priami, C., and Romanel, A. (2008). The BlenX language: a tutorial. In Formal Methods for Computational Systems Biology, ed. Bernardo, M., Degano, P., and Zavattaro, G., Berlin Heidelberg: Springer, pp. 313365.Google Scholar
de Silva, S., Ranjeewa, A. D., and Kryazhimskiy, S. (2011). The dynamics of social networks among female Asian elephants. BMC Ecology, 11(1), 17.Google Scholar
Eklöf, A., Jacob, U., Kopp, J., et al. (2013). The dimensionality of ecological networks. Ecology Letters, 16(5), 577583.Google Scholar
Faust, K. and Skvoretz, J. (2002). Comparing networks across space and time, size and species. Sociological Methodology, 32(1), 267299.Google Scholar
Gaines, M. S., Vivas, A. M., and Baker, C. L. (1979). An experimental analysis of dispersal in fluctuating vole populations: demographic parameters. Ecology, 60(4), 814828.Google Scholar
Gillespie, D. T. (1977). Exact stochastic simulation of coupled chemical reactions. Journal of Physical Chemistry, 81(25), 23402361.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(1–2), 115126.Google Scholar
Grimm, V., Berger, U., DeAngelis, D. L., et al. (2010). The ODD protocol: a review and first update. Ecological Modelling, 221(23), 27602768.Google Scholar
Hanski, I. and Zhang, D. Y. (1993). Migration, metapopulation dynamics and fugitive co-existence. Journal of Theoretical Biology, 163(4), 491504.Google Scholar
Iken, K., Brey, T., Wand, U., Voigt, J., and Junghans, P. (2001). Food web structure of the benthic community at the Porcupine Abyssal Plain (NE Atlantic): a stable isotope analysis. Progress in Oceanography, 50(1), 383405.Google Scholar
Jordán, F. and Molnár, I. (1999). Reliable flows and preferred patterns in food webs. Evolutionary Ecology Research, 1(5), 591609.Google Scholar
Jordán, F., Scotti, M., and Priami, C. (2011). Process algebra-based computational tools in ecological modelling. Ecological Complexity, 8(4), 357363.Google Scholar
Kennedy, J. and Eberhart, R. (1995) Particle swarm optimization. Proceedings of IEEE International Conference on Neural Networks, 4, 19421948.Google Scholar
Kneitel, J. M. and Chase, J. M. (2004). Trade‐offs in community ecology: linking spatial scales and species coexistence. Ecology Letters, 7(1), 6980.Google Scholar
Kokkoris, G. D., Troumbis, A. Y., and Lawton, J. H. (1999). Patterns of species interaction strength in assembled theoretical competition communities. Ecology Letters, 2(2), 7074.CrossRefGoogle Scholar
Kunz, H. and Hemelrijk, C. K. (2003). Artificial fish schools: collective effects of school size, body size, and body form. Artificial Life, 9(3), 237253.Google Scholar
Leibold, M. A., Holyoak, M., Mouquet, N., et al. (2004). The metacommunity concept: a framework for multi-scale community ecology. Ecology Letters, 7(7), 601613.Google Scholar
Levins, R. (1968). Evolution in Changing Environments: Some Theoretical Explorations. Princeton, NJ: Princeton University Press.Google Scholar
Lusseau, D., Wilson, B. E. N., Hammond, P. S., et al. (2006). Quantifying the influence of sociality on population structure in bottlenose dolphins. Journal of Animal Ecology, 75(1), 1424.Google Scholar
Marinho-Filho, J. S. (1991). The coexistence of two frugivorous bat species and the phenology of their food plants in Brazil. Journal of Tropical Ecology, 7(1), 5967.Google Scholar
Michels, E., Cottenie, K., Neys, L., and De Meester, L. (2001). Zooplankton on the move: first results on the quantification of dispersal of zooplankton in a set of interconnected ponds. Hydrobiologia, 442(1), 117126.Google Scholar
Mittermayr, A., Hansen, T., and Sommer, U. (2014a). Simultaneous analysis of δ13C, δ15N and δ34S ratios uncovers food web relationships and the trophic importance of epiphytes in an eelgrass, Zostera marina community. Marine Ecology Progress Series, 497, 93103.Google Scholar
Mittermayr, A., Fox, S. E., and Sommer, U. (2014b). Temporal variation in stable isotope composition (δ13C, δ15N and δ34S) of a temperate Zostera marina food web. Marine Ecology Progress Series, 505, 95105.Google Scholar
Murray, A. G. and Parslow, J. S. (1999). The analysis of alternative formulations in a simple model of a coastal ecosystem. Ecological Modelling, 119(2), 149166.Google Scholar
Newman, T. J., Ferdy, J. B., and Quince, C. (2004). Extinction times and moment closure in the stochastic logistic process. Theoretical Population Biology, 65(2), 115126.Google Scholar
Pacala, S. W., Gordon, D. M., and Godfray, H. C. J. (1996). Effects of social group size on information transfer and task allocation. Evolutionary Ecology, 10(2), 127165.Google Scholar
Roy, S. and Chattopadhyay, J. (2007). Towards a resolution of “the paradox of the plankton”: a brief overview of the proposed mechanisms. Ecological Complexity, 4(1), 2633.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(2), 160182.Google Scholar
Scotti, M., Hartvig, M., Winemiller, K. O., et al. (2017). Trait-based and process-oriented modeling in ecological network dynamics. In Adaptive Food Webs: Stability and Transitions of Real and Model Ecosystems, ed. Moore, J. C., de Ruiter, P. C., McCann, K. S., and Wolters, V., Cambridge, UK: Cambridge University Press, pp. 228256.Google Scholar
Stander, P. E. (1994). Cooperative hunting in lions: the role of the individual. Behavioral Ecology and Sociobiology, 29(6), 445454.Google Scholar
Sullivan, K. A. (1984). The advantages of social foraging in downy woodpeckers. Animal Behaviour, 32(1), 1622.Google Scholar
Tirelli, T. and Pessani, D. (2011). Importance of feature selection in decision-tree and artificial-neural-network ecological applications. Alburnus alburnus alborella: a practical example. Ecological Informatics, 6(5), 309315.Google Scholar
van Nouhuys, S. and Hanski, I. (2005). Metacommunities of butterflies, their host plants, and their parasitoids. In Spatial Dynamics and Ecological Communities, ed. Holyoak, M., Leibold, M. A., and Holt, R. D., Chicago: University of Chicago Press, pp. 99121.Google Scholar

References

Andersen, K. H. and Pedersen, M. (2010). Damped trophic cascades driven by fishing in model marine ecosystems. Proceeding of the Royal Society B: Biological Sciences, 277(1682), 795802.Google Scholar
Andersen, K. and Ursin, E. (1977). A multispecies extension to the Beverton and Holt theory of fishing, with accounts of phosphorus circulation and primary production. Meddelelser fra Danmarks Fiskeri- og Havundersøgelser, 7, 319435.Google Scholar
Arnold, S. J. (1972). Species densities of predators and their prey. American Naturalist, 106(948), 220236.Google Scholar
Blanchard, J. L., Andersen, K. H., Scott, F., et al. (2014). Evaluating targets and trade-offs among fisheries and conservation objectives using a multispecies size spectrum model. Journal of Applied Ecology, 51(3), 612622.Google Scholar
Cahill, A. E., Aiello-Lammens, M. E., Fisher-Reid, M. C., et al. (2013). How does climate change cause extinction? Proceeding of the Royal Society B: Biological Sciences, 280(1750), 20121890.Google Scholar
Farcas, A. and Rossberg, A. G. (2016). Maximum sustainable yield from interacting fish stocks in an uncertain world: two policy choices and underlying trade-offs. ICES Journal of Marine Science, 73(10), 24992508.Google Scholar
Flora, C., David, N., Mathias, G., and Jean-Christophe, P. (2011). Structural sensitivity of biological models revisited. Journal of Theoretical Biology, 283(1), 8291.Google Scholar
Golub, G. H. and Van Loan, C. F. (1996). Matrix Computations, third edition. Baltimore: Johns Hopkins University Press.Google Scholar
Guckenheimer, J. (2007). Bifurcation. Scholarpedia, 2(6), 1517.Google Scholar
Ives, A. R. and Carpenter, S. R. (2007). Stability and diversity of ecosystems. Science, 317, 5862.Google Scholar
Jeffries, M. J. and Lawton, J. H. (1985). Predator–prey ratios in communities of freshwater invertebrates: the role of enemy free space. Freshwater Biology, 15(1), 105112.Google Scholar
Jennings, S., Warr, K. J., and Mackinson, S. (2002). Use of size-based production and stable isotope analyses to predict trophic transfer efficiencies and predator–prey body mass ratios in food webs. Marine Ecology Progress Series, 240, 1120.Google Scholar
Jeschke, J. M. and Strayer, D. L. (2005). Invasion success of vertebrates in Europe and North America. Proceedings of the National Academy of Sciences of the United States of America, 102(28), 71987202.Google Scholar
MacArthur, R. and Levins, R. (1964). Competition, habitat selection, and character displacement in a patchy environment. Proceedings of the National Academy of Sciences of the United States of America, 51(6), 1207.Google Scholar
May, R. M. (1973). Stability and Complexity in Model Ecosystems. Princeton, NJ: Princeton University Press.Google Scholar
Menge, B. A. (1995). Indirect effects in marine rocky intertidal interaction webs: patterns and importance. Ecological Monographs, 65(1), 2174.Google Scholar
Meszéna, G., Gyllenberg, M., Pásztor, L., and Metz, J. A. J. (2006). Competitive exclusion and limiting similarity: a unified theory. Theoretical Population Biology, 69, 6887.Google Scholar
Novak, M., Wootton, J. T., Doak, D. F., et al. (2011). Predicting community responses to perturbations in the face of imperfect knowledge and network complexity. Ecology, 92(4), 836846.Google Scholar
Paine, R. T. (1980). Food webs: linkage, interaction strength and community infrastructure. Journal of Animal Ecology, 49, 667685.Google Scholar
Pugh, C. and Peixoto, M. M. (2008). Structural stability. Scholarpedia, 3(9), 4008.Google Scholar
Rescigno, A. and Richardson, I. W. (1965). On the competitive exclusion principle. Bulletin of Mathematical Biology, 27, 8589.Google Scholar
Rohr, R. P., Saavedra, S., and Bascompte, J. (2014). On the structural stability of mutualistic systems. Science, 345(6195), 1253497.Google Scholar
Rossberg, A. G. (2013). Food Webs and Biodiversity: Foundations, Models, Data. Oxford, UK: Wiley.Google Scholar
Santos de Araújo, W. (2011). Can host plant richness be used as a surrogate for galling insect diversity? Tropical Conservation Science, 4(4), 420427.Google Scholar
Sommers, H. J., Crisanti, A., Sompolinsky, H., and Stein, Y. (1988). Spectrum of large random asymmetric matrices. Physical Review Letters, 60(19), 18951898.Google Scholar
Sousa, W. P. (1984). Intertidal mosaics: patch size, propagule availability, and spatially variable patterns of succession. Ecology, 65, 19181935.Google Scholar
Speirs, D. C., Guirey, E. J., Gurney, W. S. C., and Heath, M. R. (2010). A length-structured partial ecosystem model for cod in the North Sea. Fisheries Research, 106(3), 474494.Google Scholar
Warren, P. H. and Gaston, K. J. (1992). Predator–prey ratios: a special case of a general pattern? Philosophical Transactions of the Royal Society B: Biological Sciences, 338(1284), 113130.Google Scholar
Wootton, J. T. (1994). The nature and consequences of indirect effects in ecological communities. Annual Review of Ecology and Systematics, 25, 443466.Google Scholar
Wright, M. G. and Samways, M. J. (1998). Insect species richness tracking plant species richness in a diverse flora: gall-insects in the Cape Floristic Region, South Africa. Oecologia, 115(3), 427433.Google Scholar
Yodzis, P. (1988). The indeterminacy of ecological interactions as perceived through perturbation experiments. Ecology, 69(2), 508515.Google Scholar
Yodzis, P. (1998). Local trophodynamics and the interaction of marine mammals and fisheries in the Benguela ecosystem. Journal of Animal Ecology, 67(4), 635658.Google Scholar

References

Allesina, S. and Pascual, M. (2008). Network structure, predator–prey modules, and stability in large food webs. Theoretical Ecology, 1, 5564.CrossRefGoogle Scholar
Allesina, S. and Tang, S. (2012). Stability criteria for complex ecosystems. Nature, 483, 205208.Google Scholar
Anagnostakis, S. L. (1987). Chestnut blight: the classical problem of an introduced pathogen. Mycologia, 79(1), 2337.Google Scholar
Anderson, R. M. and May, R. M. (1979). Population biology of infectious diseases: Part I. Nature, 280, 361.Google Scholar
Baer, G. M. (1991). The Natural History of Rabies. Boca Raton, FL: CRC Press.Google Scholar
Bakker, T. C., Mazzi, D., and Zala, S. (1997). Parasite-induced changes in behavior and color make Gammarus pulex more prone to fish predation. Ecology, 78, 10981104.Google Scholar
Diekmann, O., Heesterbeek, H., and Britton, T. (2013). Mathematical Tools for Understanding Infectious Disease Dynamics, Princeton, NJ: Princeton University Press.Google Scholar
Dobson, A., Lafferty, K. D., Kuris, A. M., Hechinger, R. F., and Jetz, W. (2008). Homage to Linnaeus: How many parasites? How many hosts? Proceedings of the National Academy of Sciences of the United States of America, 105, 1148211489.Google Scholar
Dunne, J. A., Lafferty, K. D., Dobson, A. P., et al. (2013). Parasites affect food web structure primarily through increased diversity and complexity. PLoS Biology, 11, e1001579.Google Scholar
Faust, E. C. (1975). The Leishmania parasite of man. In Animal Agent and Vector of Human Disease, ed. Faust, E. C., Beaver, P. C., and Jung, R. C., Philadelphia: Lea & Febiger, pp. 3345.Google Scholar
Fenton, A. and Rands, S. (2006). The impact of parasite manipulation and predator foraging behavior on predator–prey communities. Ecology, 87, 28322841.Google Scholar
Fontaine, C., Guimarães, P. R., Kéfi, S., et al. (2011). The ecological and evolutionary implications of merging different types of networks. Ecology Letters, 14, 11701181.Google Scholar
Godfray, H. C. J. (1994). Parasitoids: Behavioral and Evolutionary Ecology, Princeton, NJ: Princeton University Press.Google Scholar
Gooderham, K. and Schulte-Hostedde, A. (2011). Macroparasitism influences reproductive success in red squirrels (Tamiasciurus hudsonicus). Behavioral Ecology, 22, 11951200.Google Scholar
Hawlena, H., Khokhlova, I., Abramsky, Z., and Krasnov, B. (2006). Age, intensity of infestation by flea parasites and body mass loss in a rodent host. Parasitology, 133, 187193.Google Scholar
Hechinger, R. F., Lafferty, K. D., and Kuris, A. M. (2008). Diversity increases biomass production for trematode parasites in snails. Proceedings of the Royal Society B: Biological Sciences, 275, 27072714. DOI 10.1098/rspb.2008.0875 [doi].Google Scholar
Heesterbeek, H., Anderson, R. M., Andreasen, V., et al. (2015). Modeling infectious disease dynamics in the complex landscape of global health. Science, 347(6227), aaa4339. DOI 10.1126/science.aaa4339 [doi].Google Scholar
Huxham, M., Raffaelli, D., and Pike, A. (1995). Parasites and food web patterns. Journal of Animal Ecology, 64(2), 168176. DOI 10.2307/5752.Google Scholar
Johnson, P. T., Dobson, A., Lafferty, K. D., et al. (2010). When parasites become prey: ecological and epidemiological significance of eating parasites. Trends in Ecology and Evolution, 25, 362371. DOI 10.1016/j.tree.2010.01.005.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
Kuris, A. M. (1974). Trophic interactions: similarity of parasitic castrators to parasitoids. Quarterly Review of Biology, 49(2), 129148.Google Scholar
Kuris, A. and Lafferty, K. (2000). Parasite–host modeling meets reality: adaptive peaks and their ecological attributes. In Evolutionary Biology of Host–Parasite Relationships: Theory Meets Reality, ed. Poulin, R. and Skorping Morand, S., New York: Elsevier Science, pp. 926.Google Scholar
Kuris, A. M., Hechinger, R. F., Shaw, J. C., et al. (2008). Ecosystem energetic implications of parasite and free-living biomass in three estuaries. Nature, 454, 515518. DOI 10.1038/nature06970.Google Scholar
Lafferty, K. D. (1992) Foraging on prey that are modified by parasites. American Naturalist, 140, 854867.Google Scholar
Lafferty, K. D. (2013). Parasites in marine food webs. Bulletin of Marine Science, 89, 123134.Google Scholar
Lafferty, K. D. and Kuris, A. M. (2002). Trophic strategies, animal diversity and body size. Trends in Ecology and Evolution, 17, 507513.Google Scholar
Lafferty, K. D. and Kuris, A. M. (2009). Parasitic castration: the evolution and ecology of body snatchers. Trends in Parasitology, 25, 564572.Google Scholar
Lafferty, K. D., Dobson, A. P., and Kuris, A. M. (2006). Parasites dominate food web links. Proceedings of the National Academy of Sciences of the United States of America, 103, 1121111216. DOI 10.1073/pnas.0604755103.Google Scholar
Levin, S. A., Carpenter, S. R., Godfray, H. C. J., et al. (2009). The Princeton Guide to Ecology. Princeton, NJ: Princeton University Press.Google Scholar
Marcogliese, D. J. and Cone, D. K. (1997). Food webs: a plea for parasites. Trends in Ecology and Evolution, 12, 320325.Google Scholar
May, R. M. (1972). Will a large complex system be stable? Nature, 238, 413414.Google Scholar
May, R. M. (1973). Complexity and Stability in Model Ecosystems. Princeton, NJ: Princeton University Press.Google Scholar
McCallum, H., Harvell, D., and Dobson, A. (2003). Rates of spread of marine pathogens. Ecology Letters, 6, 10621067.Google Scholar
McCallum, H. I., Kuris, A., Harvell, C. D., et al. (2004). Does terrestrial epidemiology apply to marine systems? Trends in Ecology and Evolution, 19, 585591.Google Scholar
McCann, K. S. (2011). Food Webs. Princeton, NJ: Princeton University Press.Google Scholar
McQuaid, C. F. and Britton, N. F. (2014). Parasite species richness and its effect on persistence in food webs. Journal of Theoretical Biology, 364, 377382.Google Scholar
Moore, J. (2002). Parasites and the Behavior of Animals. Oxford, UK: Oxford University Press.Google Scholar
Moore, J. C. and de Ruiter, P. C. (2012). Energetic Food Webs: An Analysis of Real and Model Ecosystems. Oxford, UK: Oxford University Press.Google Scholar
Moret, Y. and Schmid-Hempel, P. (2000). Survival for immunity: the price of immune system activation for bumblebee workers. Science, 290, 11661168. DOI 8972 [pii].Google Scholar
Mougi, A. and Kondoh, M. (2012). Diversity of interaction types and ecological community stability. Science, 337, 349351. DOI 10.1126/science.1220529 [doi].Google Scholar
Mougi, A. and Kondoh, M. (2014). Adaptation in a hybrid world with multiple interaction types: a new mechanism for species coexistence. Ecological Research, 29, 113119.Google Scholar
Munson, L., Terio, K. A., Kock, R., et al. (2008). Climate extremes promote fatal co-infections during canine distemper epidemics in African lions. PLoS One, 3, e2545. DOI 10.1371/journal.pone.0002545.Google Scholar
Neuhaus, P. (2003). Parasite removal and its impact on litter size and body condition in Columbian ground squirrels (Spermophilus columbianus). Proceedings of the Royal Society B: Biological Sciences, 270(Suppl 2), S213S215. DOI 10.1098/rsbl.2003.0073 [doi].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. DOI 10.1126/science.1068326.Google Scholar
Nijhof, A., Cludts, S., Fisscher, O., and Laan, A. (2003). Measuring the implementation of codes of conduct: an assessment method based on a process approach of the responsible organisation. Journal of Business Ethics, 45, 6578.Google Scholar
Odum, E. P. (1971). Fundamentals of Ecology, 3rd edn. Philadelphia, PA: Saunders.Google Scholar
Packer, C., Holt, R. D., Hudson, P. J., Lafferty, K. D., and Dobson, A. P. (2003). Keeping the herds healthy and alert: implications of predator control for infectious disease. Ecology Letters, 6, 797802.Google Scholar
Pimm, S. (1982). Food Webs. London, UK: Chapman and Hall.Google Scholar
Poulin, R. (1994). Meta-analysis of parasite-induced behavioural changes. Animal Behavior, 48, 137146.Google Scholar
Poulin, R. (1999). The functional importance of parasites in animal communities: many roles at many levels? International Journal of Parasitology, 29, 903914.Google Scholar
Raveh, A., Kotler, B. P., Abramsky, Z., and Krasnov, B. R. (2011). Driven to distraction: detecting the hidden costs of flea parasitism through foraging behaviour in gerbils. Ecology Letters, 14, 4751.Google Scholar
Rip, J. and McCann, K. (2011). Cross-ecosystem differences in stability and the principle of energy flux. Ecology Letters, 14, 733740.Google Scholar
Roberts, M. and Heesterbeek, J. A. P. (2012). Characterizing the next-generation matrix and basic reproduction number in ecological epidemiology. Journal of Mathematical Biology, 66(4), 120.Google Scholar
Rossiter, W. (2013) Current opinions: zeros in host–parasite food webs: are they real? International Journal for Parasitology: Parasites and Wildlife, 2, 228234.Google Scholar
Sato, T., Egusa, T., Fukushima, K., et al. (2012). Nematomorph parasites indirectly alter the food web and ecosystem function of streams through behavioural manipulation of their cricket hosts. Ecology Letters, 15, 786793. DOI 10.1111/j.1461–0248.2012.01798.x; 10.1111/j.1461–0248.2012.01798.x.Google Scholar
Sauve, A., Fontaine, C., and Thébault, E. (2014). Structure–stability relationships in networks combining mutualistic and antagonistic interactions. Oikos, 123, 378384.Google Scholar
Schall, J. J. (1992). Parasite-mediated competition in Anolis lizards. Oecologia, 92, 5864.Google Scholar
Selaković, S., de Ruiter, P. C., and Heesterbeek, H. (2014). Infectious disease agents mediate interaction in food webs. Proceedings of the Royal Society B: Biological Sciences, 281(1777), 20132709. DOI 10.1098/rspb.2013.2709.Google Scholar
Sukhdeo, M. V. (2012). Where are the parasites in food webs? Parasites and Vectors, 5, 239. DOI 10.1186/1756–3305-5–239 [doi].Google Scholar
Suttle, C. A. (2007). Marine viruses: major players in the global ecosystem. Nature Reviews Microbiology, 5, 801812.Google Scholar
Thebault, E. and Fontaine, C. (2010). Stability of ecological communities and the architecture of mutualistic and trophic networks. Science, 329, 853856. DOI 10.1126/science.1188321 [doi].Google Scholar
Thieltges, D. W., Amundsen, P., Hechinger, R. F., et al. (2013). Parasites as prey in aquatic food webs: implications for predator infection and parasite transmission. Oikos, 122, 14731482.Google Scholar
Thompson, R. M., Mouritsen, K. N., and Poulin, R. (2004). Importance of parasites and their life cycle characteristics in determining the structure of a large marine food web. Journal of Animal Ecology, 74, 7785.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
×