Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-24T04:50:57.230Z Has data issue: false hasContentIssue false

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. 105 - 286
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

Abrams, P. A. (2000). The evolution of predator–prey interactions: theory and evidence. Annual Review of Ecology and Systematics, 31, 79105.Google Scholar
Abrams, P. A. (2006). The prerequisites for and likelihood of generalist-specialist coexistence. American Naturalist, 167, 329342.Google Scholar
Abrams, P. A. (2010). Quantitative descriptions of resource choice in ecological models. Population Ecology, 52, 4758.CrossRefGoogle Scholar
Abrams, P. A. and Matsuda, H. (1997). Prey adaptation as a cause of predator–prey cycles. Evolution, 51, 17421750.Google Scholar
Bauer, B., Vos, M., Klauschies, T., and Gaedke, U. (2014). Diversity, functional similarity and top–down control drive synchronization and the reliability of ecosystem function. American Naturalist, 183, 394409.Google Scholar
Becks, L., Ellner, S. P., Jones, L. E., and Hairston, N. G. (2010). Reduction of adaptive genetic diversity radically alters eco-evolutionary community dynamics. Ecology Letters, 13, 989997.CrossRefGoogle ScholarPubMed
Boit, A., Martinez, N. D., Williams, R. J., and Gaedke, U. (2012). Mechanistic theory and modeling of complex food web dynamics in Lake Constance. Ecology Letters, 15, 594602.Google Scholar
Bolnick, D. I., Amarasekare, P., Araujo, M. S., et al. (2011). Why intraspecific trait variation matters in community ecology. Trends in Ecology and Evolution, 26, 183192.Google Scholar
Butchart, S. H. M., Walpole, M., Collen, B., et al. (2010). Global biodiversity: indicators of recent declines. Science, 328, 11641168.Google Scholar
Chapin, F. S., Zavaleta, E. S., Eviner, V. T., et al. (2000). Consequences of changing biodiversity. Nature, 405, 234242.Google Scholar
Cortez, M. H. (2011). Comparing the qualitatively different effects rapidly evolving and rapidly induced defences have on predator–prey interactions. Ecology Letters, 14, 202209.Google Scholar
DeWitt, T. J., Sih, A., and Wilson, D. S. (1998). Costs and limits of phenotypic plasticity. Trends in Ecology and Evolution, 13, 7781.Google Scholar
Ellner, S. P. and Becks, L. (2011). Rapid prey evolution and the dynamics of two-predator food webs. Theoretical Ecology, 4, 133152.Google Scholar
Fussmann, G. F., Loreau, M., and Abrams, P. A. (2007). Eco-evolutionary dynamics of communities and ecosystems. Functional Ecology, 21, 465477.Google Scholar
Göthlich, L. and Oschlies, A. (2012). Phytoplankton niche generation by interspecific stoichiometric variation. Global Biogeochemical Cycles, 26. doi:10.1029/2011GB004042.Google Scholar
Gross, T. and Blasius, B. (2008). Adaptive coevolutionary networks: a review. Journal of the Royal Society Interface, 5, 259271.Google Scholar
Hooper, D. U., Chapin, F. S., Ewel, J. J., et al. (2005). Effects of biodiversity on ecosystem functioning: a consensus of current knowledge. Ecological Monographs, 75, 335.Google Scholar
Kasada, M., Yamamichi, M., and Yoshida, T. (2014). Form of an evolutionary tradeoff affects eco-evolutionary dynamics in a predator–prey system. Proceedings of the National Academy of Sciences of the United States of America, 111, 1603516040.CrossRefGoogle Scholar
Lande, R. (1976). Natural-selection and random genetic drift in phenotypic evolution. Evolution, 30, 314334.Google Scholar
Miner, B. E., De Meester, L., Pfrender, M. E., Lampert, W., and Hairston, N. G. Jr. (2012). Linking genes to communities and ecosystems: Daphnia as an ecogenomic model. Proceeding of the Royal Society B: Biological Sciences, 279, 18731882.Google Scholar
Mooij, W. M., Trolle, D., Jeppesen, E., et al. (2010). Challenges and opportunities for integrating lake ecosystem modelling approaches. Aquatic Ecology, 44, 633667. doi:10.1007/s10452-010–9339-3 ER.Google Scholar
Mougi, A. (2012). Unusual predator–prey dynamics under reciprocal phenotypic plasticity. Journal of Theoretical Biology, 305, 96102.Google Scholar
Murdoch, W. W. (1973). Functional response of predators. Journal of Applied Ecology, 10, 335342.Google Scholar
Naeem, S., Bunker, D. E., Hector, A., et al. (eds.) (2009). Biodiversity, Ecosystem Functioning, and Human Wellbeing: An Ecological and Economic Perspective. Oxford, UK: Oxford University Press.Google Scholar
Norberg, J., Urban, M. C., Vellend, M., Klausmeier, C. A., and Loeuille, N. (2012). Eco-evolutionary responses of biodiversity to climate change. Nature Climate Change, 2, 747751.Google Scholar
Tirok, K. and Gaedke, U. (2010). Internally driven alternation of functional traits in a multi-species predator–prey system. Ecology, 91, 17481762.Google Scholar
Tirok, K., Bauer, B., Wirtz, K., and Gaedke, U. (2011). Predator–prey dynamics driven by feedback between functionally diverse trophic levels. PLoS ONE, 6, e27357. doi:10.1371/journal.pone.0027357.Google Scholar
Urban, M. C., Leibold, M. A., Amarasekare, P., et al. (2008). The evolutionary ecology of metacommunities. Trends in Ecology and Evolution, 23(6), 311317.Google Scholar
van Velzen, E. and Etienne, R. S. (2013). The evolution and coexistence of generalist and specialist herbivores under between-plant competition. Theoretical Ecology, 6, 8798.Google Scholar
van Velzen, E. and Etienne, R. S. (2015). The importance of ecological costs for the evolution of plant defense against herbivory. Journal of Theoretical Biology, 372, 8999.Google Scholar
Vos, M., Kooi, B. W., DeAngelis, D. L., and Mooij, W. M. (2004). Inducible defences and the paradox of enrichment. Oikos, 105, 471480.Google Scholar
Yamamichi, M., Yoshida, T., and Sasaki, A. (2011). Comparing the effects of rapid evolution and phenotypic plasticity on predator–prey dynamics. American Naturalist, 178, 287304.CrossRefGoogle ScholarPubMed
Yoshida, T., Jones, L. E., Ellner, S. P., Fussmann, G. F., and Hairston, N. G. Jr. (2003). Rapid evolution drives ecological dynamics in a predator–prey system. Nature, 424, 303306.Google Scholar

References

Abrams, P. A. (2011). Simple life-history omnivory: responses to enrichment and harvesting in systems with intraguild predation. American Naturalist, 178(3), 305319.Google Scholar
Abrams, P. A. and Walters, C. J. (1996). Invulnerable prey and the paradox of enrichment. Ecology, 77(4), 11251133.Google Scholar
Andersen, K. H. and Pedersen, M. (2010). Damped trophic cascades driven by fishing in model marine ecosystems. Proceedings of the Royal Society B: Biological Sciences, 277, 795802.Google Scholar
Baum, J. K. and Worm, B. (2009). Cascading top–down effects of changing oceanic predator abundances. Journal of Animal Ecology, 78(4), 699714.Google Scholar
Benoît, E. and Rochet, M. J. (2004). A continuous model of biomass size spectra governed by predation and the effects of fishing on them. Journal of Theoretical Biology, 226(1), 921.Google Scholar
Berlow, E. L. (1999). Strong effects of weak interactions in ecological communities. Nature, 398(6725), 330.Google Scholar
Blanchard, J. L., Jennings, S., Law, R., et al. (2009). Size-spectra dynamics from stochastic predation and growth of individuals. Journal of Animal Ecology, 78(1), 270280.Google Scholar
Blanchard, J. L., Law, R., Castle, M. D., and Jennings, S. (2011). Coupled energy pathways and the resilience of size-structured food webs. Theoretical Ecology, 4(3), 289300.Google Scholar
Boerlijst, M. C., Oudman, T., and de Roos, A. M. (2013). Catastrophic collapse can occur without early warning: examples of silent catastrophes in structured ecological models. PloS One, 8(4).Google Scholar
Bolnick, D. I., Amarasekare, P., Araújo, M. S., et al. (2011). Why intraspecific trait variation matters in community ecology. Trends in Ecology & Evolution, 26(4), 183192.CrossRefGoogle ScholarPubMed
Brose, U., Williams, R. J., and Martinez, N. D. (2006a). Allometric scaling enhances stability in complex food webs. Ecology Letters, 9(11), 12281236.Google Scholar
Brose, U., Jonsson, T., Berlow, E., et al. (2006b). Consumer-resource body-size relationships in natural food webs. Ecology, 87(10), 24112417.Google Scholar
Brose, U., Blanchard, J. L., Eklöf, A., et al. (2016). Predicting the consequences of species loss using size-structured biodiversity approaches. Biological Reviews. doi:10.1111/brv.12250.Google Scholar
Brunel, T. and Piet, G. J. (2013). Is age structure a relevant criterion for the health of fish stocks? ICES Journal of Marine Science, 70, 270283.Google Scholar
Byström, P. and Andersson, J. (2005). Size‐dependent foraging capacities and intercohort competition in an ontogenetic omnivore (Arctic char). Oikos, 3, 523536.Google Scholar
Byström, P. and Garcia-Berthou, E. (1999). Density dependent growth and size specific competitive interactions in young fish. Oikos, 86, 217232.Google Scholar
Casini, M., Lövgren, J., Hjelm, J., et al. (2008). Multi-level trophic cascades in a heavily exploited open marine ecosystem. Proceedings of the Royal Society B: Biological Sciences, 275, 17931801.CrossRefGoogle Scholar
Casini, M., Hjelm, J., Molinero, J.-C., et al. (2009). Trophic cascades promote threshold-like shifts in pelagic marine ecosystems. Proceedings of the National Academy of Sciences of the United States of America, 106, 197202.Google Scholar
Caskenette, A. L. and McCann, K. S. (2017). Biomass reallocation between juveniles and adults mediates food web stability by distributing energy away from strong interactions. PLoS ONE, 12(1), e0170725. doi:10.1371/journal.pone.0170725.Google Scholar
Charnov, E. L. (2001). Reproductive efficiencies in the evolution of life histories. Evolutionary Ecology Research, 3(7), 873876.Google Scholar
Chesson, P. (2000). Mechanisms of maintenance of species diversity. Annual Review of Ecology and Systematics, 31(2000), 343358.Google Scholar
Chesson, P. and Kuang, J. J. (2008). Competition and biodiversity in spatially structured habitats. Nature, 456(7219), 235238.Google Scholar
Claessen, D., de Roos, A. M., and Persson, L. (2004). Population dynamic theory of size-dependent cannibalism. Proceedings of the Royal Society B: Biological Sciences, 271(1537), 333340.Google Scholar
de Roos, A. M. and Persson, L. (2001). Physiologically structured models: from versatile technique to ecological theory. Oikos, 94(1), 5171.Google Scholar
de Roos, A. M. and Persson, L. (2003). Competition in size-structured populations: mechanisms inducing cohort formation and population cycles. Theoretical Population Biology, 63(1), 116.Google Scholar
de Roos, A. M. and Persson, L. (2013). Population and Community Ecology of Ontogenetic Development. Princeton, NJ: Princeton University Press.Google Scholar
de Roos, A. M., Persson, L., and McCauley, E. (2003a). The influence of size-dependent life-history traits on the structure and dynamics of populations and communities. Ecology Letters, 6(5), 473487.Google Scholar
de Roos, A. M., Persson, L., and Thieme, H. R. (2003b). Emergent Allee effects in top predators feeding on structured prey populations. Proceedings of the Royal Society B: Biological Sciences, 270(1515), 611618.CrossRefGoogle ScholarPubMed
de Roos, A. M., Schellekens, T., van Kooten, T., van de Wolfshaar, K. E., Claessen, D., and Persson, L. (2007). Food-dependent growth leads to overcompensation in stage-specific biomass when mortality increases: the influence of maturation versus reproduction regulation. American Naturalist, 170(3), E59E76.Google Scholar
de Roos, A. M., Schellekens, T., van Kooten, T., and Persson, L. (2008a). Stage-specific predator species help each other to persist while competing for a single prey. Proceedings of the National Academy of Sciences of the United States of America, 105(37), 1393013935.Google Scholar
de Roos, A. M., Schellekens, T., van Kooten, T., van de Wolfshaar, K. E., Claessen, D., and Persson, L. (2008b). Simplifying a physiologically structured population model to a stage-structured biomass model. Theoretical Population Biology, 73(1), 4762.Google Scholar
Digel, C., Curtsdotter, A., Riede, J., Klarner, B., and Brose, U. (2014). Unravelling the complex structure of forest soil food webs: higher omnivory and more trophic levels. Oikos, 123(10), 11571172.Google Scholar
Eklöf, A., Jacob, U., Kopp, J., et al. (2013). The dimensionality of ecological networks. Ecology Letters, 16(5), 577583.Google Scholar
English, K. K., Edgell, T. C., Bocking, R. C., Link, M., and Raborn, S. (2011). Fraser River sockeye fisheries and fisheries management and comparison with Bristol Bay sockeye fisheries. LGL Ltd. Cohen Commission Technical Report 7, 190.Google Scholar
Estes, J. A., Terborgh, J., Brashares, J. S., et al. (2011). Trophic downgrading of planet Earth. Science, 333(6040), 301306.Google Scholar
Gårdmark, A., Casini, M., Huss, M., et al. (2014). Regime shifts in exploited marine food-webs: detecting mechanisms underlying alternative stable states using size-structured community dynamics theory. Philosophical Transactions of the Royal Society B: Biological Sciences, 370, DOI: 10.1098/rstb.2013.0262.Google Scholar
Gilljam, D., Thierry, A., Edwards, F. K., et al. (2011). Seeing double: size-based and taxonomic views of food web structure. Advances in Ecological Research, 45, 67133.CrossRefGoogle Scholar
Glazier, D. S. (2005). Beyond the “3/4-power law”: variation in the intra- and interspecific scaling of metabolic rate in animals. Biological Reviews of the Cambridge Philosophical Society, 80(4), 611662.Google Scholar
Guill, C. (2009). Alternative dynamical states in stage-structured consumer populations. Theoretical Population Biology, 76(3), 168178.Google Scholar
Gurney, W. and Nisbet, R. (1985). Fluctuation periodicity, generation separation, and the expression of larval competition. Theoretical Population Biology, 180, 150180.Google Scholar
Hartvig, M. (2011). Food Web Ecology: Individual Life-Histories and Ecological Processes Shape Complex Communities. Ph.D. thesis, Department of Biology, Lund University, Sweden.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
Hartvig, M., Andersen, K. H., and Beyer, J. E. (2011). Food web framework for size-structured populations. Journal of Theoretical Biology, 272(1), 113122.Google Scholar
Hastings, A. (1983). Age-dependent predation is not a simple process. 1. Continuous –time models. Theoretical Population Biology, 23(3), 347362.Google Scholar
Heckmann, L., Drossel, B., Brose, U., and Guill, C. (2012). Interactive effects of body-size structure and adaptive foraging on food-web stability. Ecology Letters, 15(3), 243250.Google Scholar
Hidalgo, M., Rouyer, T., Molinero, J. C., et al. (2011). Synergistic effects of fishing-induced demographic changes and climate variation on fish population dynamics. Marine Ecology Progress Series, 2011(426), 112.CrossRefGoogle Scholar
Hin, V., Schellekens, T., Persson, L., and de Roos, A. M. (2011). Coexistence of predator and prey in intraguild predation systems with ontogenetic niche shifts. American Naturalist, 178(6), 701714.Google Scholar
Huss, M. and Nilsson, K. A. (2011). Experimental evidence for emergent facilitation: promoting the existence of an invertebrate predator by killing its prey. Journal of Animal Ecology, 80(3), 615621.Google Scholar
Hutchings, J. and Myers, R. (1994). What can be learned from the collapse of a renewable resource? Atlantic cod, Gadus morhua, of Newfoundland and Labrador. Canadian Journal of Fisheries and Aquatic Sciences, 51, 21262146.Google Scholar
Jennings, S., Pinnegar, J. K., Polunin, N. V. C., and Boon, T. W. (2001). Weak cross-species relationships between body size and trophic level belie powerful size-based trophic structuring in fish communities. Ecology Letters, 70(6), 934944.Google Scholar
Kartascheff, B., Heckmann, L., Drossel, B., and Guill, C. (2010). Why allometric scaling enhances stability in food web models. Theoretical Ecology, 3(3), 195208.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(4), 291300.Google Scholar
Kitchell, J. F., Stewart, D. J., and Weininger, D. (1977). Applications of a bioenergetics model to yellow perch (Perca flavescens) and walleye (Stizostedion vitreum vitreum). Journal of the Fisheries Research Board of Canada, 34(10), 19221935.Google Scholar
Kooijman, S. A. L. M. (2000). Dynamic Energy and Mass Budgets in Biological Systems. Cambridge, UK: Cambridge University Press.Google Scholar
Law, R., Plank, M. J., James, A., and Blanchard, J. L. (2009). Size-spectra dynamics from stochastic predation and growth of individuals. Ecology, 90(3), 802811.CrossRefGoogle ScholarPubMed
Magnússon, K. G. (1999). Destabilizing effect of cannibalism on a structured predator–prey system. Mathematical Biosciences, 155, 6175.Google Scholar
Martinez, N. (1991). Artifacts or attributes? Effects of resolution on the Little Rock Lake food web. Ecological Monographs, 61(4), 367392.Google Scholar
Matsuda, H. and Abrams, P. A. (2006). Maximal yields from multispecies fisheries systems: rules for systems with multiple trophic levels. Ecological Applications, 16(1), 225237.Google Scholar
May, R. M. (1972). Will a large complex system be stable? Nature, 238, 413414.Google Scholar
May, R. M. (2006). Network structure and the biology of populations. Trends in Ecology and Evolution, 21(7), 394399.Google Scholar
McCann, K., Hastings, A., and Huxel, G. (1998). Weak trophic interactions and the balance of nature. Nature, 395, 794798.Google Scholar
Metz, J. A. J. and Diekmann, O. (1986). The Dynamics of Physiologically Structured Populations. Lecture Notes in Biomathematics, vol. 68. Amsterdam: Springer-Verlag.Google Scholar
Mougi, A. and Kondoh, M. (2012). Diversity of interaction types and ecological community stability. Science, 349(2012), 337.Google Scholar
Murdoch, W. W., Kendall, B. E., Nisbet, R. M., et al. (2002). Single-species models for many-species food webs. Nature, 417(6888), 541543.CrossRefGoogle ScholarPubMed
Murphy, L. F. and Smith, S. J. (1990). Optimal harvesting of an age-structured population. Journal of Mathematical Biology, 29, 7790.Google Scholar
Mylius, S. D., Klumpers, K., de Roos, A. M., and Persson, L. (2001). Impact of intraguild predation and stage structure on simple communities along a productivity gradient. American Naturalist, 158(3), 259276.Google Scholar
Nakazawa, T. (2011). Ontogenetic niche shift, food-web coupling, and alternative stable states. Theoretical Ecology, 4(4), 479494.Google Scholar
Newton, P. F. (1997). Stand density management diagrams: review of their development and utility in stand-level management planning. Forest Ecology and Management, 98, 251265.Google Scholar
Ohlberger, J., Langangen, Ø., Edeline, E., et al. (2011). Stage-specific biomass overcompensation by juveniles in response to increased adult mortality in a wild fish population. Ecology, 92(12), 21752182.Google Scholar
Oksanen, L., Fretwell, S. D., Arruda, J., and Niemela, P. (1981). Exploitation ecosystems in gradients of primary productivity. American Naturalist, 118(2), 240261.Google Scholar
Persson, L. (1999). Trophic cascades: abiding heterogeneity and the trophic level concept at the end of the road. Oikos, 85(3), 385397.CrossRefGoogle Scholar
Persson, L., de Roos, A. M., Claessen, D., et al. (2003). Gigantic cannibals driving a whole-lake trophic cascade. Proceedings of the National Academy of Sciences of the United States of America, 100(7), 40354039.Google Scholar
Persson, L., Amundsen, P.-A., de Roos, A. M., Klemetsen, A., Knudsen, R., and Primicerio, R. (2007). Culling prey promotes predator recovery: alternative states in a whole-lake experiment. Science, 316(5832), 17431746.Google Scholar
Persson, L., van Leeuwen, A., and de Roos, A. M. (2014). The ecological foundation for ecosystem-based management of fisheries: mechanistic linkages between the individual-, population-, and community-level dynamics. ICES Journal of Marine Science, 71(8), 22682280.Google Scholar
Peters, R. H. (1983). The Ecological Implications of Body Size. New York: Cambridge University Press.Google Scholar
Pikitch, E., Santora, C., Babcock, E., et al. (2004). Policy forum: ecosystem-based fishery management. Science, 305, 346347.Google Scholar
Pimm, S. L. and Rice, J. C. (1987). The dynamics of multispecies, multi-life-stage models of aquatic food webs. Theoretical Population Biology, 32(3), 303325.Google Scholar
Pimm, S., Lawton, J., and Cohen, J. (1991). Food web patterns and their consequences. Nature, 350, 669674.Google Scholar
Plank, M. J. and Law, R. (2012). Size-spectra dynamics from stochastic predation and growth of individuals. Theoretical Ecology, 5(4), 465480.Google Scholar
Polis, G. (1984). Age structure component of niche width and intraspecific resource partitioning: can age groups function as ecological species? American Naturalist, 123(4), 541564.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, 1605, 29232934.Google Scholar
Rijnsdorp, A. D. (1993). Fisheries as a large-scale experiment on life-history evolution: disentangling phenotypic and genetic effects in changes in maturation and reproduction of North Sea plaice, Pleuronectes platessa L. Oecologia, 96(3), 391401.Google Scholar
Rochet, M.-J. and Benoit, E. (2012). Fishing destabilizes the biomass flow in the marine size spectrum. Proceedings of the Royal Society B: Biological Sciences, 279(1727), 284292.Google Scholar
Rooney, N., McCann, K., Gellner, G., and Moore, J. C. (2006). Structural asymmetry and the stability of diverse food webs. Nature, 442(7100), 265269.CrossRefGoogle ScholarPubMed
Rosenzweig, M. and MacArthur, R. (1963). Graphical representation and stability conditions of predator–prey interactions. American Naturalist, 97(895), 209.Google Scholar
Rossberg, A. G., Brännström, A., and Dieckmann, U. (2010a). Food-web structure in low- and high-dimensional trophic niche spaces. Journal of the Royal Society Interface, 7(53), 17351743.Google Scholar
Rossberg, A. G., Brännström, Å., and Dieckmann, U. (2010b). How trophic interaction strength depends on traits. Theoretical Ecology, 3(1), 1324.Google Scholar
Rudolf, V. H. W. and Lafferty, K. D. (2011). Stage structure alters how complexity affects stability of ecological networks. Ecology Letters, 14(1), 7579.Google Scholar
Rudolf, V. H. W. and Rasmussen, N. L. (2013a). Ontogenetic functional diversity: size structure of a keystone predator drives functioning of a complex ecosystem. Ecology, 94(5), 10461056.Google Scholar
Rudolf, V. H. W. and Rasmussen, N. L. (2013b). Population structure determines functional differences among species and ecosystem processes. Nature Communications, 4, 2318.Google Scholar
Scheffer, M., Carpenter, S., Foley, J. A., Folke, C., and Walker, B. (2001). Catastrophic shifts in ecosystems. Nature, 413(6856), 591596.Google Scholar
Scheffer, M., Bascompte, J., Brock, W. A., et al. (2009). Early-warning signals for critical transitions. Nature, 461(7260), 5359.Google Scholar
Schellekens, T., de Roos, A. M., and Persson, L. (2010). Ontogenetic diet shifts result in niche partitioning between two consumer species irrespective of competitive abilities. American Naturalist, 176(5), 625637.Google Scholar
Schreiber, S. and Rudolf, V. H. W. (2008). Crossing habitat boundaries: coupling dynamics of ecosystems through complex life cycles. Ecology Letters, 11(6), 576587.Google Scholar
Silvert, W. and Platt, T. (1980). Dynamic energy-flow model of the particle size distribution in pelagic ecosystems. In Evolution and Ecology of Zooplankton Communities, ed. Kerfoot, W. C., Illanover, NH: University Press of New England, pp. 754763.Google Scholar
Stearns, S. (1989). Trade-offs in life-history evolution. Functional Ecology, 3(3), 259268.Google Scholar
Tilman, D. (1994). Competition and biodiversity in spatially structured habitats. Ecology, 75(1), 12281236.Google Scholar
Ursin, E. (1973). On the prey size preferences of cod and dab. Meddelelser fra Danmarks Fiskeri-og Havundersgelser, 7, 8598.Google Scholar
van de Wolfshaar, K., de Roos, A., and Persson, L. (2006). Size‐dependent interactions inhibit coexistence in intraguild predation systems with life‐history omnivory. American Naturalist, 168(1), 6275.Google Scholar
van de Wolfshaar, K. E., HilleRisLambers, R., and Gårdmark, A. (2011). Effect of habitat productivity and exploitation on populations with complex life cycles. Marine Ecology Progress Series, 438, 175184.Google Scholar
van Kooten, T., Persson, L., and de Roos, A. M. (2007). Size-dependent mortality induces life-history changes mediated through population dynamical feedbacks. American Naturalist, 170, 258270.Google Scholar
van Leeuwen, A., de Roos, A. M., and Persson, L. (2008). How cod shapes its world. Journal of Sea Research, 60(1–2), 89104.Google Scholar
Walters, C. and Kitchell, J. F. (2001). Cultivation/depensation effects on juvenile survival and recruitment: implications for the theory of fishing. Canadian Journal of Fisheries and Aquatic Sciences, 58(1), 3950.Google Scholar
Werner, E. and Gilliam, J. (1984). The ontogenetic niche and species interactions in size-structured populations. Annual Review of Ecology and Systematics, 15, 393425.Google Scholar
Woodward, G., Blanchard, J., Lauridsen, R. B., et al. (2010). Individual-based food webs: species identity, body size and sampling effects. Advances in Ecological Research, 43, 211266.Google Scholar
Yodzis, P. (1981). The stability of real ecosystems. Nature, 289, 674676.Google Scholar
Zhang, L., Thygesen, U. H., Knudsen, K., and Andersen, K. H. (2013). Trait diversity promotes stability of community dynamics. Theoretical Ecology, 6(1), 5769.Google Scholar

References

Abrams, P. A. and Matsuda, H. (1997). Prey adaptation as a cause of predator–prey cycles. Evolution, 51, 17421750.Google Scholar
Bauer, B., Vos, M., Klauschies, T., and Gaedke, U. (2014). Diversity, functional similarity and top–down control drive synchronization and the reliability of ecosystem function. American Naturalist, 183, 394409.Google Scholar
Becks, L., Ellner, S. P., Jones, L. E., and Hairston, N. G. (2010). Reduction of adaptive genetic diversity radically alters eco-evolutionary community dynamics. Ecology Letters, 13, 989997.Google Scholar
Becks, L., Ellner, S. P., Jones, L. E., and Hairston, N. G. Jr. (2012). The functional genomics of an eco-evolutionary feedback loop, linking gene expression, trait evolution, and community dynamics. Ecology Letters, 15, 492501.Google Scholar
Binzer, A., Guill, C., Brose, U., and Rall, B. C. (2012). The dynamics of food chains under climate change and nutrient enrichment. Philosophical Transactions of the Royal Society B: Biological Sciences, 367, 29352944.Google Scholar
Boit, A., Martinez, N. D., Williams, R. J., and Gaedke, U. (2012). Mechanistic theory and modeling of complex food web dynamics in Lake Constance. Ecology Letters, 15, 594602.Google Scholar
Bolnick, D. I., Amarasekare, P., Araujo, M. S., et al. (2011). Why intraspecific trait variation matters in community ecology. Trends in Ecology and Evolution, 26, 183192.Google Scholar
Brose, U., Jonsson, T., Berlow, E. L., et al. (2006a). Consumer-resource body-size relationships in natural food webs. Ecology, 87, 24112417.Google Scholar
Brose, U., Williams, R. J., and Martinez, N. D. (2006b). Allometric scaling enhances stability in complex food webs. Ecology Letters, 9, 12281236.Google Scholar
Conti, L., Schmidt-Kloibe, A., Grenouillet, G., and Graf, W. (2014). A trait-based approach to assess the vulnerability of European aquatic insects to climate change. Hydrobiologia, 721(1), 297315.Google Scholar
Cortez, M. H. (2011). Comparing the qualitatively different effects rapidly evolving and rapidly induced defences have on predator–prey interactions. Ecology Letters, 14, 202209.Google Scholar
Cortez, M. H. and Ellner, S. P. (2010). Understanding rapid evolution in predator–prey interactions using the theory of fast–slow dynamical systems. American Naturalist, 176, 109127.Google Scholar
De Roos, A. M. and Persson, L. (2001). Physiologically structured models: from versatile technique to ecological theory. Oikos, 94(1), 5171.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
De Roos, A. M., Schellekens, T., van Kooten, T., et al. (2007). Food-dependent growth leads to overcompensation in stage-specific biomass when mortality increases: the influence of maturation versus reproduction regulation. American Naturalist, 170, 5976.Google Scholar
De Roos, A. M., Schellekens, T., van Kooten, T., et al. (2008). Simplifying a physiologically structured population model to a stage-structured biomass model. Theoretical Population Biology, 73, 4762.Google Scholar
De Roos, A. M. and Persson, P. (2013). Population and Community Ecology of Ontogenetic Development. Princeton, NJ: Princeton University Press.Google Scholar
Dieckmann, U. and Law, R. (1996). The dynamical theory of coevolution: a derivation from stochastic ecological processes. Journal of Mathematical Biology, 34, 579612.Google Scholar
Digel, C., Curtsdotter, A., Riede, J., Klarner, B., and Brose, U. (2014). Unravelling the complex structure of forest soil food webs: higher omnivore and more trophic levels. Oikos, 123, 11571172.Google Scholar
dos Santos, F. A. S., Johst, K., and Grimm, V. (2011). Neutral communities may lead to decreasing diversity–disturbance relationships: insights from a generic simulation model. Ecology Letters, 14, 653660.Google Scholar
Duffy, J. E., Cardinale, B. J., France, K. E., et al. (2007). The functional role of biodiversity in ecosystems: incorporating trophic complexity. Ecology Letters, 10, 522538.Google Scholar
Edwards, K. F., Klausmeier, C. A., and Litchman, E. (2013a). A three-way tradeoff maintains functional diversity under variable resource supply. American Naturalist, 182, 786800.Google Scholar
Edwards, K. F., Klausmeier, C. A., and Litchman, E. (2013b). Functional traits predict variation in phytoplankton community structure across lakes of the United States. Ecology, 94, 16261635.Google Scholar
Feio, M. J. and Doledec, S. (2012). Integration of invertebrate traits into predictive models for indirect assessment of stream functional integrity: a case study in Portugal. Ecological Indicators, 15, 236247.Google Scholar
Gallagher, R. V., Hughes, L., and Leishman, M. R. (2013). Species loss and gain in communities under future climate change: consequences for functional diversity. Ecography, 36(5), 531540.Google Scholar
Gray, D. K. and Arnott, S. E. (2011). Does dispersal limitation impact the recovery of zooplankton communities damaged by a regional stressor? Ecological Applications, 21, 12411256.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
Grover, J. P. (1991). Resource competition in a variable environment: phytoplankton growing according to the Variable–Internal–Stores model. American Naturalist, 138, 811835.Google Scholar
Guill, C. (2009). Alternative dynamical states in stage-structured consumer populations. Theoretical Population Biology, 76, 168178.Google Scholar
Heckmann, L., Drossel, B., Brose, U., and Guill, C. (2012). Interactive effects of body-size structure and adaptive foraging on food-web stability. Ecology Letters, 15, 243250.Google Scholar
Hillebrand, H. and Matthiessen, B. (2009). Biodiversity in a complex world: consolidation and progress in functional biodiversity research. Ecology Letters, 12, 14051419.Google Scholar
Hooper, D. U., Chapin, F. S., Ewel, J. J., et al. (2005). Effects of biodiversity on ecosystem functioning: a consensus of current knowledge. Ecological Monographs, 75, 335.Google Scholar
Janse, J. H., De Senerpont Domis, L. N., Scheffer, M., et al. (2008). Critical phosphorus loading of different types of shallow lakes and the consequences for management estimated with the ecosystem model PCLake. Limnologica – Ecology and Management of Inland Waters, 38(3), 203219.Google Scholar
Janse, J. H., Scheffer, M., Lijklema, L., et al. (2010). Estimating the critical phosphorus loading of shallow lakes with the ecosystem model PCLake: sensitivity, calibration and uncertainty. Ecological Modelling, 221(4), 654665. doi:10.1016/j.ecolmodel.2009.07.023 ER.Google Scholar
Jones, L. E. and Ellner, S. P. (2007). Effects of rapid prey evolution on predator–prey cycles. Journal of Mathematical Biology, 55, 541573.Google Scholar
Jones, L. E., Becks, L., Ellner, S. P., et al. (2009). Rapid contemporary evolution and clonal food web dynamics. Philosophical Transactions of the Royal Society B: Biological Sciences, 364, 15791591.Google Scholar
Jørgensen, S. E. (1994). Models as instruments for combination of ecological theory and environmental practice. Ecological Modelling, 75, 520.Google Scholar
Kalinkat, G., Schneider, F. D., Digel, C., et al. (2013). Body masses, functional responses and predator–prey stability. Ecology Letters, 16, 11261134.Google Scholar
Kooijman, S. A. L. M. (2010). Dynamic Energy Budget Theory for Metabolic Organisation. Cambridge, UK: Cambridge University Press.Google Scholar
Litchman, E., Klausmeier, C. A., Schofield, O. M., and Falkowski, P. G. (2007). The role of functional traits and trade-offs in structuring phytoplankton communities: scaling from cellular to ecosystem level. Ecology Letters, 10, 11701181.CrossRefGoogle ScholarPubMed
May, F., Grimm, V., and Jeltsch, F. (2009). Reversed effects of grazing on plant diversity: the role of below-ground competition and size symmetry. Oikos, 118, 18301843.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
Merico, A., Brandt, G., Smith, S. L., and Oliver, M. (2014). Sustaining diversity in trait-based models of phytoplankton communities. Frontiers in Ecology and Evolution, 2, 59, 18.Google Scholar
Metz, J. A. J. and Diekmann, O. (1986). The Dynamics of Physiologically Structured Populations. Springer-Verlag.Google Scholar
Mooij, W. M., Trolle, D., Jeppesen, E., et al. (2010). Challenges and opportunities for integrating lake ecosystem modelling approaches. Aquatic Ecology, 44(3), 633667. doi:10.1007/s10452-010–9339-3 ER.Google Scholar
Mougi, A. (2012). Unusual predator–prey dynamics under reciprocal phenotypic plasticity. Journal of Theoretical Biology, 305, 96102.Google Scholar
Naeem, S. and Wright, J. P. (2003). Disentangling biodiversity effects on ecosystem functioning: deriving solutions to a seemingly insurmountable problem. Ecology Letters, 6, 567579.Google Scholar
Nakazawa, T. (2011). Ontogenetic niche shift, food-web coupling, and alternative stable states. Theoretical Ecology, 4, 479494.Google Scholar
Norberg, J. (2004). Biodiversity and ecosystem functioning: a complex adaptive systems approach. Limnology and Oceanography, 49, 12691277.Google Scholar
Petchey, O. L., Beckerman, A. P., Riede, J. O., and Warren, P. H. (2008). Size, foraging, and food-web structure. Proceeding of the National Academy of Sciences, 105, 41914196.Google Scholar
Peters, R. H. (1993). The Ecological Implications of Body Size. Cambridge University Press.Google Scholar
Ponce-Reyes, R., Nicholson, E., Baxter, P. W. J., Fuller, R. A., and Possingham, H. (2013). Extinction risk in cloud forest fragments under climate change and habitat loss. Biodiversity Research, 19, 518529.Google Scholar
Post, D. M. and Palkovacs, E. P. (2009). Eco-evolutionary feedbacks in community and ecosystem ecology: interactions between the ecological theatre and the evolutionary play. Philosophical Transactions of the Royal Society B: Biological Sciences, 364, 16291640.Google Scholar
Savage, V. M. and Norberg, J. (2007). A general multi-trait-based framework for studying the effects of biodiversity on ecosystem functioning. Journal of Theoretical Biology, 247, 213229.Google Scholar
Smith, S. L. and Yamanaka, Y. (2007). Optimization-based model of multinutrient uptake kinetics. Limnology and Oceanography, 52, 15451558.Google Scholar
Solan, M., Cardinale, B. J., Downing, A. L., et al. (2004). Extinction and ecosystem function in the marine benthos. Science, 306, 11771180.Google Scholar
Sollie, S., Janse, J. H., Mooij, W. M., Coops, H., and Verhoeven, J. T. A. (2008). The contribution of marsh zones to water quality in Dutch shallow lakes: a modeling study. Environmental Management, 42(6), 10021016. doi:10.1007/s00267-008–9121-7.Google Scholar
Sommer, U., Padisak, J., Reynolds, C. S., and Juhasz-Hagy, P. (1993). Hutchinson’s heritage: the diversity–disturbance relationship in phytoplankton. Hydrobiologia, 249, 17.Google Scholar
Sommer, U., Sommer, F., Santer, B., et al. (2003). Daphnia versus copepod impact on summer phytoplankton: functional compensation at both trophic levels. Oecologia, 135, 639647.Google Scholar
Sommer, U., Adrian, R., Domis, L. D. S., et al. (2012). Beyond the Plankton Ecology Group (PEG) model: mechanisms driving plankton succession. Annual Review of Ecology, Evolution and Systematics, 43, 429448. doi:10.1146/annurev-ecolsys-110411–160251 ER.Google Scholar
Sterner, R. W. and Elser, J. J. (2002). Ecological Stoichiometry: The Biology of Elements from Molecules to the Biosphere. Princeton, NJ: Princeton University Press.Google Scholar
Tirok, K. and Gaedke, U. (2007). Regulation of planktonic ciliate dynamics and functional composition during spring in Lake Constance. Aquatic Microbial Ecology, 49, 87100.Google Scholar
Tirok, K. and Gaedke, U. (2010). Internally driven alternation of functional traits in a multi-species predator–prey system. Ecology, 91, 17481762.Google Scholar
Tirok, K., Bauer, B., Wirtz, K., and Gaedke, U. (2011). Predator–prey dynamics driven by feedback between functionally diverse trophic levels. PLoS ONE, 6 (11), e27357. doi:10.1371/journal.pone.0027357.Google Scholar
Tomimatsu, H., Sasaki, T., Kurokawa, H., et al. (2013). Sustaining ecosystem functions in a changing world: a call for an integrated approach. Journal of Applied Ecology, 50(5), 11241130.Google Scholar
Urban, M. C., Leibold, M. A., Amarasekare, P., et al. (2008). The evolutionary ecology of metacommunities. Trends in Ecology and Evolution, 23(6) 311317.Google Scholar
van der Stap, I., Vos, M., and Mooij, W. M. (2007). Induced defenses in herbivores and plants differentially modulate a trophic cascade. Ecology, 88, 24742481.Google Scholar
Van Gerven, L. P. A., de Klein, J. J. M., Gerla, D. J., et al. (2015). Competition for light and nutrients in layered communities of aquatic plants. American Naturalist, 186(1), 7283.Google Scholar
Verschoor, A. M., Vos, M., and van der Stap, I. (2004). Inducible defences prevent strong population fluctuations in bi- and tritrophic food chains. Ecology Letters, 7, 11431148.Google Scholar
Violle, C., Navas, M.-L., Vile, D., et al. (2007). Let the concept of trait be functional! Oikos, 116, 882892.Google Scholar
Violle, C., Enquist, B. J., McGill, B. J., et al. (2012). The return of the variance: intraspecific variability in community ecology. Trends in Ecology and Evolution, 27, 244252.Google Scholar
Wagner, A. and Benndorf, J. (2007). Climate-driven warming during spring destabilises a Daphnia population: a mechanistic food-web approach. Oecologia, 151, 351364.Google Scholar
Werner, E. E. and Gilliam, J. F. (1984). The ontogenetic niche and species interactions in size-structured populations. Annual Review of Ecology and Systematics, 15, 393425.Google Scholar
Williams, R. J. and Martinez, N. D. (2000). Simple rules yield complex food webs. Nature, 404, 180183.Google Scholar
Wirtz, K. W. and Eckhardt, B. (1996). Effective variables in ecosystem models with an application to phytoplankton succession. Ecological Modelling, 92, 3353.Google Scholar
Woodward, G. and Hildrew, A. G. (2002). Body-size determinants of niche overlap and intraguild predation within a complex food web. Journal of Animal Ecology, 71, 10631074.Google Scholar
Yoshida, T., Jones, L. E., Ellner, S. P., Fussmann, G. F., and Hairston, N. G. Jr. (2003). Rapid evolution drives ecological dynamics in a predator–prey system. Nature, 424, 303306.Google Scholar
Yoshida, T., Ellner, S. P., Jones, L. E., et al. (2007). Cryptic population dynamics: rapid evolution masks trophic interactions. PLoS Biology, 5, 18681879.Google Scholar
Zhang, L., Thygesen, U. H., Knudsen, K., and Andersen, K. H. (2013). Trait diversity promotes stability of community dynamics. Theoretical Ecology, 6, 5769.Google Scholar

References

Allesina, S. and Bodini, A. (2008). Ascendency. In Encyclopedia of Ecology Vol. 1, ed. Jørgensen, S. E. and Fath, B. D., Oxford: Elsevier, pp. 254263.Google Scholar
Antonietti, R., Ferrari, I., Rossetti, G., Tarozzi, L., and Viaroli, P. (1988). Zooplankton structure in an oligotrophic mountain lake in Northern Italy. Verhandlungen des Internationalen Verein Limnologie., 23, 545552.Google Scholar
Archer, S. and Stokes, C. (2000). Stress, disturbance and change in rangeland ecosystems: rangeland desertification. Advances in Vegetation Science, 19, 1738.Google Scholar
Bertani, I., Primicerio, R., and Rossetti, G. (2016). Extreme climatic event triggers a lake regime shift that propagates across multiple trophic levels. Ecosystems, 19, 1631.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.Google Scholar
Clements, F. E. (1936). Nature and structure of the climax. Journal of Ecology, 24, 252284.Google Scholar
Drury, W. H. and Nisbet, I. C. T. (1973). Succession. Journal of the Arnold Arboretum, 54, 331368.Google Scholar
Kerfoot, W. C. (ed.) (1980). The Evolution and Ecology of Zooplankton Communities. Hanover, NH: University Press of New England.Google Scholar
Kerfoot, W. C. and Sih, A. (eds.) (1987). Predation: Direct and Indirect Impacts on Aquatic Communities. Hanover, NH: University Press of New England.Google Scholar
Latham, L. G. and Scully, E. P. (2002). Quantifying constraint to assess development in ecological networks. Ecological Modelling, 154, 2544.Google Scholar
Lynch, M. and Shapiro, J. (1981). Predation, enrichment, and phytoplankton community structure. Limnology and Oceanography, 26, 86102.Google Scholar
MacArthur, R. (1955). Fluctuation of animal populations and a measure of community stability. Ecology, 36, 533536.Google Scholar
MacMahon, J. A. (1980). Ecosystems over time: succession and other types of change. In Forests: Fresh Perspectives from Ecosystem Analysis, ed. Waring, R. H. and Corvalis, O. R., Oregon State University Press, pp. 2758.Google Scholar
Mageau, M. T., Costanza, R., and Ulanowicz, R. E. (1995). The development and initial testing of a quantitative assessment of ecosystem health. Ecosystem Health, 1, 201213.Google Scholar
Mageau, M. T., Costanza, R., and Ulanowicz, R. E. (1998). Quantifying the trends expected in developing ecosystems. Ecological Modelling, 112, 122.Google Scholar
Noble, I. R. and Slatyer, R. O. (1977). Post-fire succession of plants in Mediterranean ecosystems. In Proceedings of the symposium on the environmental consequences of fire and fuel management in Mediterranean climate ecosystems. USDA Forest Service General Technical Report. WO-3, pp. 2736.Google Scholar
Odum, E. P. (1969). The strategy of ecosystem development. Science, 164, 262270.Google Scholar
Odum, E. P. (1985). Trends expected in stressed ecosystems. BioScience, 35, 419422.Google Scholar
Oksanen, L. (1991). Trophic levels or trophic dynamics: a consensus emerging? Trends in Ecology and Evolution, 6(2), 5860.Google Scholar
Paris, G., Rossetti, G., Cattadori, M., and Giordani, G. (1995). Phytoplankton–zooplankton interactions in a small mountain lake (Lake Scuro, Parma Appennines): results from enclosure experiments. Proceedings of the Italian Society of Ecology, 18, 147150.Google Scholar
Rossetti, G., Hamzah, W., and Paris, G. (1997). Zooplankton grazing activity and algal food electivity in a mountain lake. Proceedings of the Italian Society of Ecology, 18, 147150.Google Scholar
Rutledge, R. W., Basorre, B. L., and Mulholland, R. J. (1976). Ecological stability: an information theory viewpoint. Journal of Theoretical Biology, 57, 355371.Google Scholar
Schindler, D. W. (1990). Experimental perturbations of whole lakes as tests of hypotheses concerning ecosystem structure and function. Oikos, 57, 2541.Google Scholar
Sousa, W. P. (1984). Intertidal mosaics: patch size, propagule availability, and spatially variable patterns of succession. Ecology, 65, 19181935.Google Scholar
Ulanowicz, R. E. (1986). Growth and Development: Ecosystems Phenomenology. New York: Springer-Verlag.Google Scholar
Ulanowicz, R. E. (1997). Ecology: The Ascendent Perspective. New York: Columbia University Press.Google Scholar
Ulanowicz, R. E. (2001). Information theory in ecology. Computers and Chemistry, 25, 393399.Google Scholar
Ulanowicz, R. E. (2004). Quantitative methods for ecological network analysis. Computational Biology and Chemistry, 28, 321339.Google Scholar
Whittaker, R. H. (1953). A consideration of climax theory: the climax as a population and pattern. Ecological Monographs, 23, 4178.Google Scholar

References

Allesina, S. and Pascual, M. (2009). Food web models: a plea for groups. Ecology Letters, 12, 652662.Google Scholar
Allesina, S., Alonso, D., and Pascual, M. (2008). A general model for food web structure. Science, 320, 658661.Google Scholar
Bersier, L. F., Cattin, M. F., Banasek-Richter, C., Baltensperger, R., and Gabriel, J. P. (2006). Box B: Reply to Martinez and Cushing. In Ecological Networks: Linking Structure to Dynamics in Food Webs, ed. Pascual, M. and Dunne, J. A., New York: Oxford University Press, pp. 9192.Google Scholar
Brose, U., Cushing, L., Berlow, E. L., et al. (2005). Body sizes of consumers and their resources. Ecology, 86, 2545.Google Scholar
Capitán, J. A., Arenas, A., and Guimerà, R. (2013). Degree of intervality of food webs: from body-size data to models. Journal of Theoretical Biology, 334, 3544.Google Scholar
Cattin, M. F., Bersier, L. F., Banasek-Richter, C., Baltensperger, R., and Gabriel, J. P. (2004). Phylogenetic constraints and adaptation explain food-web structure. Nature, 427, 835839.Google Scholar
Cohen, J. E. (1968). Alternate derivations of a species-abundance relation. American Naturalist, 102, 165172.Google Scholar
Cohen, J. E. (1978). Food Webs and Niche Space. Princeton, NJ: Princeton University Press.Google Scholar
Cohen, J. E. and Newman, C. M. (1985). A stochastic theory of community food webs. 1. Models and aggregated data. Proceedings of the Royal Society B: Biological Sciences, 224, 421448.Google Scholar
Cohen, J. E., Briand, F., and Newman, C. M. (1990). Community Food Webs, Data and Theory. Berlin: Springer-Verlag.Google Scholar
Cohen, J. E., Jonsson, T., and Carpenter, S. R. (2003). Ecological community description using the food web, species abundance, and body size. Proceedings of the National Academy of Sciences of the United States of America, 100, 17811786.Google Scholar
Drossel, B., Higgs, P. G., and McKane, A. J. (2001). The influence of predator–prey population dynamics on the long-term evolution of food web structure. Journal of Theoretical Biology, 208, 91107.Google Scholar
Eklöf, A., Jacob, U., Kopp, J., et al. (2013). The dimensionality of ecological networks. Ecology Letters, 16, 577583.Google Scholar
Fournier, T., Rohr, R. P., Scherer, H., Mazza, C., and Bersier, L. F. (2009). How to estimate the niche values in a food web? The 94th ESA Annual Meeting (August 2–7, 2009), abstract: http://eco.confex.com/eco/2009/techprogram/P20894.HTM.Google Scholar
Freckleton, R. P., Harvey, P. H., and Pagel, M. (2002). Phylogenetic analysis and comparative data: a test and review of evidence. American Naturalist, 160, 712726.Google Scholar
Grafen, A. (1989). The phylogenetic regression. Philosophical Transactions of the Royal Society B: Biological Sciences, 326, 119157.Google Scholar
Gravel, D., Poisot, T., Albouy, C., Velez, L., and Mouillot, D. (2013). Inferring food web structure from predator–prey body size relationships. Methods in Ecology and Evolution, 4, 10831090.Google Scholar
Harper-Smith, S., Berlow, E. L., Knapp, R. A., Williams, R. J., and Martinez, N. D. (2005). Communicating ecology through food webs: visualizing and quantifying the effects of stocking alpine lakes with trout. In Dynamic Food Webs: Multispecies Assemblages, Ecosystem Development, and Environmental Change, ed. de Ruiter, P C., Wolters, V., and Moore, J. C., Elsevier Academic Press, pp. 407423.Google Scholar
Hoff, P. D. (2009). Multiplicative latent factors models for description and prediction of social networks. Computational and Mathematicals Organization, 15, 261272.Google Scholar
Jonsson, T., Cohen, J. E., and Carpenter, S. R. (2005). Food webs, body size, and species abundance in ecological community description. In Advances in Ecological Research, Vol. 36, ed. Caswell, H., Elsevier Academic Press, pp. 184.Google Scholar
Kolaczyk, E. D. (2009). Statistical Analysis of Network Data. New York: Springer.Google Scholar
Legendre, P. and Legendre, L. (1998). Numerical Ecology. Amsterdam: Elsevier.Google Scholar
Naisbit, R. E., Rohr, R. P., Rossberg, A. G., Kehrli, P., and Bersier, L.-F. (2012). Phylogeny versus body size as determinants of food web structure. Proceedings of the Royal Society B: Biological Sciences, 279, 32913297.Google Scholar
Pellissier, L., Rohr, R. P., Ndiribe, C., et al. (2013). Combining food web and species distribution models for improved community projections. Ecology and Evolution, 3, 45724583.Google Scholar
Petchey, O. L., Beckerman, A. P., Riede, J. O., and Warren, P. H. (2008). Size, foraging, and food web structure. Proceedings of the National Academy of Sciences of the United States of America, 105, 41914196.Google Scholar
Pinnegar, J. K., Trenkel, V. M., Tidd, A. N., Dawson, W. A., and Du Buit, M. H. (2003). Does diet in Celtic Sea fishes reflect prey availability? In Annual Symposium of the Fisheries Society of the British Isles, Norwich, pp. 197212.Google Scholar
Price, P. W. (2003). Macroevolutionary Theory on Macroecological Patterns. Cambridge, UK: Cambridge University Press.Google Scholar
Rohr, R. P., Scherer, H., Kehrli, P., Mazza, C., and Bersier, L. F. (2010). Modeling food webs: exploring unexplained structure using latent traits. American Naturalist, 173, 170177.Google Scholar
Rohr, R. P., Naisbit, R. E., Mazza, C., and Bersier, L. F. (2016). Matching-centrality decomposition and the forecasting of new links in networks. Proceedings of the Royal Society B: Biological Sciences, 283, 20152702.Google Scholar
Ross, R. (1911). Some quantitative studies in epidemiology. Nature, 87, 466467.Google Scholar
Rossberg, A. G. (2013). Food Webs and Biodiversity: Foundations, Models, Data. Wiley.Google Scholar
Rossberg, A. G., Matsuda, H., Amemiya, T., and Itoh, K. (2006). Food webs: experts consuming families of experts. Journal of Theoretical Biology, 241, 552563.Google Scholar
Rossberg, A. G., Brannstrom, A., and Dieckmann, U. (2010). How trophic interaction strength depends on traits. Theoretical Ecology, 3, 1324.Google Scholar
Stouffer, D. B., Camacho, J., Guimera, R., Ng, C. A., and Nunes Amaral, L. A. (2005). Quantitative patterns in the structure of model and empirical food webs. Ecology, 86, 13011311.Google Scholar
Stouffer, D. B., Camacho, J., and Amaral, L. A. N. (2006). A robust measure of food web intervality. Proceedings of the National Academy of Sciences of the United States of America, 103, 1901519020.Google Scholar
Warren, P. H. (1989). Spatial and temporal variation in the structure of a freshwater food web. Oikos, 55, 299311.Google Scholar
Williams, R. J. and Martinez, N. D. (2000). Simple rules yield complex food webs. Nature, 404, 180183.Google Scholar
Williams, R. J. and Martinez, N. D. (2008). Success and its limits among structural models of complex food webs. Journal of Animal Ecology, 77, 512519.Google Scholar
Williams, R. J. and Purves, D. W. (2011). The probabilistic niche model reveals substantial variation in the niche structure of empirical food webs. Ecology, 92, 18491857.Google Scholar
Williams, R. J., Anandanadesan, A., and Purves, D. (2010). The probabilistic niche model reveals the niche structure and role of body size in a complex food web. PLoS ONE, 5, e12092.Google Scholar
Woodward, G., Speirs, D. C., and Hildrew, A. G. (2005). Quantification and resolution of a complex, size-structured food web. In Advances in Ecological Research, Vol. 36, ed. Caswell, H., Elsevier Academic Press, pp. 85135.Google Scholar

References

Alford, D. V. (2011). Plant Pests. London: Harper Collins Publishers.Google Scholar
Alford, D. V., Nilsson, C., and Ulber, B. (2003). Insect pests of oilseed rape crops. In Biocontrol of Oilseed Rape Pests, ed. Alford, D. V., Oxford: Blackwell Publishing, pp. 941.Google Scholar
Almeida-Neto, M. and Ulrich, W. (2011). A straightforward computational approach for measuring nestedness using quantitative matrices. Environmental Modelling and Software, 26, 173178.Google Scholar
Almeida-Neto, M., Guimaraes, P., Guimaraes, P. R. Jr., Loyola, R. D., and Ulrich, W. (2008). A consistent metric for nestedness analysis in ecological systems: reconciling concept and measurement. Oikos, 117, 12271239.Google Scholar
Angelsen, A. (2010). Policies for reduced deforestation and their impact on agricultural production. Proceedings of the National Academy of Sciences of the United States of America, 107, 1963919644.Google Scholar
Attwood, S. J., Maron, M., House, A. P. N., and Zammit, C. (2008). Do arthropod assemblages display globally consistent responses to intensified agricultural land use and management? Global Ecology and Biogeography, 17, 585599.Google Scholar
Barbosa, P. (1998). Agroecosystems and conservation biological control. In Conservation Biological Control, ed. Barbosa, P., San Diego: Academic Press, pp. 3954.Google Scholar
Bardwell, C. J. and Averill, A. L. (1997). Spiders and their prey in Massachusetts cranberry bogs. Journal of Arachnology, 25, 3141.Google Scholar
Bascompte, J., Jordano, P., Melian, 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
Bilsing, S. W. (1920). Quantitative studies in the food of spiders. Ohio Journal of Science, 20, 215260.Google Scholar
Birkhofer, K. and Wolters, V. (2012). The global relationship between climate, net primary production and the diet of spiders. Global Ecology and Biogeography, 21, 100108.Google Scholar
Birkhofer, K., Scheu, S., and Wise, D. H. (2007). Small-scale spatial pattern of web-building spiders (Araneae) in Alfalfa: relationship to disturbance from cutting, prey availability, and intraguild interactions. Environmental Entomology, 36, 801810.Google Scholar
Birkhofer, K., Entling, M., and Lubin, Y. (2013). Agroecology: trait composition, spatial relationships, trophic interactions. In Spider Research in the 21st Century: Trends and Perspectives, ed. Penney, D., Manchester: Siri Scientific Press, pp. 200229.Google Scholar
Birkhofer, K., Arvidsson, F., Ehlers, D., et al. (2015) Landscape complexity and organic farming independently affect the biological control of hemipteran pests and yields in spring barley. Landscape Ecology, 31, 567579. DOI: 10.1007/s10980-015–0263-8.Google Scholar
Blitzer, E. J., Dormann, C. F., Holzschuh, A., et al. (2012). Spillover of functionally important organisms between managed and natural habitats. Agriculture Ecosystems & Environment, 146, 3443.Google Scholar
Bluthgen, N., Menzel, F., and Bluthgen, N. (2006). Measuring specialization in species interaction networks. BMC Ecology, 6, 9.Google Scholar
Bohan, D. A. and Woodward, G. (2013). Editorial commentary: the potential for network approaches to improve knowledge, understanding, and prediction of the structure and functioning of agricultural systems. Advances in Ecological Research, 49, xiiixviii.Google Scholar
Bohan, D. A., Raybould, A., Mulder, C., et al. (2013). Networking agroecology: integrating the diversity of agroecosystem interactions. Advances in Ecological Research, 49, 167.Google Scholar
Dąbrowska-Prot, E., Łuczak, J., and Tarwid, K. (1968). Prey and predator density and their reactions in the process of mosquitoes reduction by spiders in field experiments. Ekologia Polska, 16, 773819.Google Scholar
DeFries, R. S., Rudel, T., Uriarte, M., and Hansen, M. (2010). Deforestation driven by urban population growth and agricultural trade in the twenty-first century. Nature Geoscience, 3, 178181.Google Scholar
Diehl, E., Mader, V. L., Wolters, V., and Birkhofer, K. (2013). Management intensity and vegetation complexity affect web-building spiders and their prey. Oecologia, 173, 579589.Google Scholar
Dormann, C. F. and Strauss, R. (2014). A method for detecting modules in quantitative bipartite networks. Methods in Ecology and Evolution, 5, 9098.Google Scholar
Dormann, C. F., Gruber, B., and Fründ, J. (2008). Introducing the bipartite package: analysing ecological networks. R News, 8, 811.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., Santa Fe, NM: Santa Fe Institute Studies in the Sciences of Complexity, 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
Estes, J. A., Terborgh, J., Brashares, J. S., et al. (2011). Trophic downgrading of planet Earth. Science, 333, 301306.Google Scholar
Finlay-Doney, M. and Walter, G. H. (2012). The conceptual and practical implications of interpreting diet breadth mechanistically in generalist predatory insects. Biological Journal of the Linnean Society, 107, 737763.Google Scholar
Gray, C., Baird, D. J., Baumgartner, S., et al. (2014). Ecological networks: the missing links in biomonitoring science. Journal of Applied Ecology, 51, 14441449.Google Scholar
Guseinov, E. F. (2005). Natural prey of the jumping spider Salticus tricinctus (Araneae, Salticidae). Bulletin of the British Arachnological Society, 13, 130132.Google Scholar
Guseinov, E. F., Cerveira, A. M., and Jackson, R. R. (2004). The predatory strategy, natural diet, and life cycle of Cyrba algerina, an araneophagic jumping spider (Salticidae: Spartaeinae) from Azerbaijan. New Zealand Journal of Zoology, 31, 291303.Google Scholar
Harwood, J. D., Sunderland, K. D., and Symondson, W. O. C. (2001). Living where the food is: web location by linyphiid spiders in relation to prey availability in winter wheat. Journal of Applied Ecology, 38, 8899.Google Scholar
Havlík, P., Schneider, U. A., Schmid, E., et al. (2011). Global land-use implications of first and second generation biofuel targets. Energy Policy, 39, 56905702.Google Scholar
Heleno, R., Devoto, M., and Pocock, M. (2012). Connectance of species interaction networks and conservation value: is it any good to be well connected? Ecological Indicators, 14, 710.Google Scholar
Hines, J., van der Putten, W. H., de Deyn, G., et al. (2015). Towards an integration of biodiversity–ecosystem functioning and food web theory to advance the understanding of connections between multiple ecosystem functions and service provisioning. Advances in Ecological Research, 253, 161199.Google Scholar
Huseynov, E. F. o. (2005). Natural prey of the jumping spider Menemerus taeniatus (Araneae: Salticidae). European Journal of Entomology, 102, 797799.Google Scholar
Huseynov, E. F. o. (2006a). The prey of the lynx spider Oxyopes globifer (Araneae, Oxyopidae) associated with a semidesert dwarf shrub in Azerbaijan. Journal of Arachnology, 34, 422426.Google Scholar
Huseynov, E. F. o. (2006b). Natural prey of the jumping spider Heliophanus dunini (Araneae: Salticidae) associated with Eryngium plants. Bulletin of the British Arachnological Society, 13, 293296.Google Scholar
Huseynov, E. F. o. (2007a). Natural prey of the crab spider Thomisus onustus (Araneae: Thomisidae), an extremely powerful predator of insects. Journal of Natural History, 41, 23412349.Google Scholar
Huseynov, E. F. o. (2007b). Natural prey of the lynx spider Oxyopes lineatus (Araneae: Oxyopidae). Entomologica Fennica, 18, 144148.Google Scholar
Huseynov, E. F. o. (2007c). Natural prey of the crab spider Runcinia grammica (Araneae: Thomisidae) on Eryngium plants. Bulletin of the British Arachnological Society, 14, 9396.Google Scholar
Huseynov, E. F. o. (2008). Natural prey of the jumping spider Philaeus chrysops (Araneae: Salticidae) in different types of microhabitat. Bulletin of the British Arachnological Society, 14, 262268.Google Scholar
Huseynov, E. F., Cross, F. R., and Jackson, R. R. (2005). Natural diet and prey-choice behaviour of Aelurillus muganicus (Araneae: Salticidae), a myrmecophagic jumping spider from Azerbaijan. Journal of Zoology, 267, 159165.Google Scholar
Ings, T. C., Bascompte, M. J. M., Blüthgen, N., et al. (2009). Ecological networks: beyond food webs. Journal of Animal Ecology, 78, 253269.Google Scholar
Ives, A. R., Cardinale, B. J., and Snyder, W. E. (2005). A synthesis of subdisciplines: predator–prey interactions, and biodiversity and ecosystem functioning. Ecology Letters, 8, 102116.Google Scholar
Jensen, K., Mayntz, D., Toft, S., Raubenheimer, D., and Simpson, S. J. (2011). Prey nutrient composition has different effects on Pardosa wolf spiders with dissimilar life histories. Oecologia, 165, 577583.Google Scholar
Kankaanpää, S. and Carter, T. R. (2004). An Overview of Forest Policies Affecting Land Use in Europe. The Finnish Environment 706, Helsinki: Finnish Environment Institute.Google Scholar
Kiritani, K., Kawahara, S., Sasaba, T., and Nakasuji, F. (1972). Quantitative evaluation of predation by spiders on the green rice leafhopper Nephotettix cincticeps by a sight count method. Researches on Population Ecology, 13, 187200.Google Scholar
Kondoh, M., Kato, S., and Sakato, Y. (2010). Food webs are built up with nested subwebs. Ecology, 91, 31233130.Google Scholar
Kuusk, A.-K. and Ekbom, B. (2010). Lycosid spiders and alternative food: feeding behavior and implications for biological control. Biological Control, 55, 2026.Google Scholar
Kuusk, A.-K. and Ekbom, B. (2012). Feeding habits of lycosid spiders in field habitats. Journal of Pest Science, 85, 253260.Google Scholar
Laliberté, E. and Tylianakis, J. M. (2010). Deforestation homogenizes tropical parasitoid–host networks. Ecology, 91, 17401747.Google Scholar
Lambin, E. F. and Meyfroidt, P. (2011). Global land use change, economic globalization, and the looming land scarcity. Proceedings of the National Academy of Sciences of the United States of America, 108, 34653472.Google Scholar
Landis, D. A., Wratten, S. D., and Gurr, G. M. (2000). Habitat management to conserve natural enemies of arthropod pests in agriculture. Annual Review of Entomology, 45, 175201.Google Scholar
Layer, K., Hildrew, A. G., Jenkins, G., et al. (2011). Long-term dynamics of a well-characterised food web: four decades of acidification and recovery in the Broadstone stream model system. Advances in Ecological Research, 44, 69117.Google Scholar
Lewinsohn, T. M., Prado, P. I., Jordano, P., Bascompte, J., and Olesen, J. M. (2006). Structure in plant–animal interaction assemblages. Oikos, 113, 174184.Google Scholar
Lewis, T. (1997). Pest thrips in perspective. In Thrips as Crop Pests, ed. Lewis, T., Wallingford: CABI, pp. 113.Google Scholar
Marc, P., Canard, A., and Ysnel, F. (1999). Spiders (Araneae) useful for pest limitation and bioindication. Agriculture Ecosystems & Environment, 74, 229273.Google Scholar
McKinlay, R. G. (1992). Vegetable Crop Pests. Boca Raton: CRC Press.Google Scholar
Miranda, M., Parrini, F., and Dalerum, F. (2013). A categorization of recent network approaches to analyse trophic interactions. Methods in Ecology and Evolution, 4, 897905.Google Scholar
Müller, J., Bussler, H., Gossner, M. M., Rettelbach, T., and Duelli, P. (2008). The European spruce bark beetle Ips typographus in a national park: from pest to keystone species. Biodiversity and Conservation, 17, 29793001.Google Scholar
Nentwig, W. (1982). Analyses of the prey of cribellate spiders (Araneae: Filistatidae, Dictynidae, Eresidae). Entomologische Mitteilungen aus dem zoologischen Museum Hamburg, 7, 233244.Google Scholar
Nentwig, W. (1983a). The non-filter function of orb webs in spiders. Oecologia, 58, 418420.Google Scholar
Nentwig, W. (1983b). The prey of web-building spiders compared with feeding experiments (Araneae, Araneidae, Linyphiidae, Pholcidae, Agelenidae). Oecologia, 56, 132139.Google Scholar
Nentwig, W. (1985). Prey analysis of 4 species of tropical orb-weaving spiders (Araneae, Araneidae) and a comparison with Araneids of the temperate zone. Oecologia, 66, 580594.Google Scholar
Nentwig, W. (1987). The prey of spiders. In Ecophysiology of Spiders, ed. Nentwig, W., Berlin: Springer Verlag, pp. 249263.Google Scholar
Netherer, S. and Schopf, A. (2010). Potential effects of climate change on insect herbivores in European forests: general aspects and the pine processionary moth as specific example. Forest Ecology and Management, 259, 831838.Google Scholar
Nyffeler, M. (1999). Prey selection of spiders in the field. Journal of Arachnology, 27, 317324.Google Scholar
Nyffeler, M. and Benz, G. (1978). Prey selection by web spiders Argiope bruennichi (Scop.), Araneus quadratus (Cl.), and Agelena labyrinthica (Cl.) on fallow land near Zurich, Switzerland. Revue Suisse De Zoologie, 85, 747757.Google Scholar
Nyffeler, M. and Benz, G. (1979). Overlap of the niches concerning space and prey of crab spiders (Araneae, Thomisidae) and wolf spiders (Araneae, Lycosidae) in cultivated meadows. Revue Suisse De Zoologie, 86, 855865.Google Scholar
Nyffeler, M. and Benz, G. (1981a). Field studies on the feeding ecology of spiders: observations in the region of Zurich (Switzerland). Anzeiger für Schädlingskunde Pflanzenschutz Umweltschutz, 54, 3339.Google Scholar
Nyffeler, M. and Benz, G. (1981b). Some observations on the feeding ecology of the wolf-spider Pardosa lugubris (walck). Deutsche Entomologische Zeitschrift, 28, 297300.Google Scholar
Nyffeler, M. and Benz, G. (1988a). Prey and predatory importance of micryphantid spiders in winter-wheat fields and hay meadows. Journal of Applied Entomology, 105, 190197.Google Scholar
Nyffeler, M. and Benz, G. (1988b). Feeding ecology and predatory importance of wolf spiders (Pardosa spp.) (Araneae, Lycosidae) in winter-wheat fields. Journal of Applied Entomology, 106, 123134.Google Scholar
Nyffeler, M. and Benz, G. (1988c). Prey analysis of the spider Achaearanea riparia (Blackw.) (Araneae, Theridiidae), a generalist predator in winter-wheat fields. Journal of Applied Entomology, 106, 425431.Google Scholar
Nyffeler, M. and Sterling, W. L. (1994). Comparison of the feeding niche of polyphagous insectivores (Araneae) in a Texas cotton plantation: estimates of niche breadth and overlap. Environmental Entomology, 23, 12941303.Google Scholar
Nyffeler, M. and Sunderland, K. D. (2003). Composition, abundance and pest control potential of spider communities in agroecosystems: a comparison of European and US studies. Agriculture Ecosystems and Environment, 95, 579612.Google Scholar
Nyffeler, M., Dean, D. A., and Sterling, W. L. (1986). Feeding-habits of the spiders Cyclosa turbinata (Walckenaer) (Araneae, Araneidae) and Lycosa rabida Walckenaer (Araneae, Lycosidae). Southwestern Entomologist, 11, 195201.Google Scholar
Nyffeler, M., Dean, D. A., and Sterling, W. L. (1987). Predation by green lynx spider, Peucetia viridans (Araneae, Oxyopidae), inhabiting cotton and woolly croton plants in east Texas. Environmental Entomology, 16, 355359.Google Scholar
Nyffeler, M., Dean, D. A., and Sterling, W. L. (1988). Prey records of the web-building spiders Dictyna segregata (Dictynidae), Theridion australe (Theridiidae), Tidarren haemorrhoidale (Theridiidae), and Frontinella pyramitela (Linyphiidae) in a cotton agroecosystem. Southwestern Naturalist, 33, 215218.Google Scholar
Nyffeler, M., Dean, D. A., and Sterling, W. L. (1992). Diets, feeding specialization, and predatory role of 2 lynx spiders, Oxyopes salticus and Peucetia viridans (Araneae, Oxyopidae), in a Texas cotton agroecosystem. Environmental Entomology, 21, 14571465.Google Scholar
Olesen, J. E. and Bindi, M. (2002). Consequences of climate change for European agricultural productivity, land use and policy. European Journal of Agronomy, 16, 239262.Google Scholar
Pekar, S., Coddington, J. A., and Blackledge, T. A. (2012). Evolution of stenophagy in spiders (Araneae): evidence based on the comparative analysis of spider diets. Evolution, 66, 776806.Google Scholar
Perez-De la Cruz, M., Sanchez-Soto, S., Ortiz-Garcia, C. F., Zapata-Mata, R., and De la Cruz-Perez, A. (2007). Diversity of insects captured by weaver spiders (Arachnida: Araneae) in the cocoa agroecosystem in Tabasco, Mexico. Neotropical Entomology, 36, 90101.Google Scholar
Poisot, T., Canard, E., Mouillot, D., Mouquet, N., and Gravel, D. (2012) The dissimilarity of species interaction networks. Ecology Letters, 15, 13531361.Google Scholar
Pyle, R., Bentzien, M., and Opler, P. (1981). Insect conservation. Annual Review of Entomology, 26, 233258.Google Scholar
R Development Core Team (2008). R: A Language and Environment for Statistical Computing. Vienna: R Foundation for Statistical Computing.Google Scholar
Rabbinge, R. and van Diepen, C. A. (2000). Changes in agriculture and land use in Europe. European Journal of Agronomy, 13, 8599.Google Scholar
Rand, T. A., Tylianakis, J. M., and Tscharntke, T. (2006). Spillover edge effects: the dispersal of agriculturally subsidized insect natural enemies into adjacent natural habitats. Ecology Letters, 9, 603614.Google Scholar
Rounsevell, M. D. A., Reginster, I., Araújo, M. B., et al. (2006). A coherent set of future land use change scenarios for Europe. Agriculture Ecosystems and Environment, 114, 5768.Google Scholar
Sala, O. E., Chaplin, F. S., Armesto, J. J., et al. (2000). Biodiversity: global biodiversity scenarios for the year 2100. Science, 287, 17701774.Google Scholar
Sandoval, C. P. (1994). Plasticity in web design in the spider Parawixia bistriata: a response to variable prey type. Functional Ecology, 8, 701707.Google Scholar
Schmitz, C., van Meijl, H., Kyle, P., et al. (2014). Land‐use change trajectories up to 2050: insights from a global agro‐economic model comparison. Agricultural Economics, 45, 6984.Google Scholar
Smith, H. G., Birkhofer, K., Clough, Y., et al. (2014) Beyond dispersal: the role of animal movement in modern agricultural landscapes. In Animal Movement Across Scales, ed. Hansson, L. A. and Åkesson, S., Oxford: Oxford University Press.Google Scholar
Smith, P., Gregory, P. J., Van Vuuren, D., et al. (2010). Competition for land. Philosophical Transactions of the Royal Society B: Biological Sciences, 365, 29412957.Google Scholar
Spittlehouse, D. L. and Stewart, R. B. (2004). Adaptation to climate change in forest management. Journal of Ecosystems and Management, 4, 111.Google Scholar
Stouffer, D. B., Sales-Pardo, M., Sirer, M. I., and Bascompte, J. (2012). Evolutionary conservation of species’ roles in food webs. Science, 335, 14891492.Google Scholar
Sunderland, K. D., Powell, W., and Symondson, W. O. C. (2005). Populations and communities. In Insects as Natural Enemies: A Practical Perspective, ed. Jervis, M. A., Dordrecht: Springer, pp. 299434.Google Scholar
Thebault, E. and Fontaine, C. (2008). Does asymmetric specialization differ between mutualistic and trophic networks? Oikos, 117, 555563.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, 689697.Google Scholar
Tilman, D. (1999). Global environmental impacts of agricultural expansion: the need for sustainable and efficient practices. Proceedings of the National Academy of Sciences of the United States of America, 96, 59956000.Google Scholar
Tilman, D., Fargione, J., Wolff, B., et al. (2001). Forecasting agriculturally driven global environmental change. Science, 292, 281284.Google Scholar
Tixier, P., Peyrard, N., Auberlot, J.-N., et al. (2013). Modelling interaction networks for enhanced ecosystem services in agroecosystems. Advances in Ecological Research, 49, 437480.Google Scholar
Traugott, M., Kamenova, S., Ruess, L., Seeber, J., and Plantegenest, M. (2013). Empirically characterising trophic networks: what emerging DNA-based methods, stable isotope and fatty acid analyses can offer. Advances in Ecological Research, 49, 177224.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, 202205.Google Scholar
Tylianakis, J. M., Laliberté, E., Nielsen, A., and Bascompte, J. (2010). Conservation of species interaction networks. Biological Conservation, 143, 22702279.Google Scholar
Uetz, G. W. and Hartsock, S. P. (1987). Prey selection in an orb-weaving spider Micrathena gracilis (Araneae, Araneidae). Psyche, 94, 103116.Google Scholar
Uetz, G. W., Johnson, A. D., and Schemske, D. W. (1978). Web placement, web structure, and prey capture in orb-weaving spiders. Bulletin of the British Arachnological Society, 4, 141148.Google Scholar
van Emden, H. F., and Harrington, R. (2007). Aphids as Crop Pests. Wallingford: CABI.Google Scholar
van der Putten, W. H., de Ruiter, P. C., Bezemer, T. M., et al. (2004). Trophic interactions in a changing world. Basic and Applied Ecology, 5, 487494.Google Scholar
Wise, D. H. (1993). Spiders in Ecological Webs. Cambridge, Cambridge University Press.Google Scholar
Wise, D. H. and Barata, J. L. (1983). Prey of 2 syntopic spiders with different web structures. Journal of Arachnology, 11, 271281.Google Scholar

References

Alemanno, S., Mancinelli, G., and Basset, A. (2007). Effects of invertebrate patch use behaviour and detritus quality on reed leaf decomposition in aquatic systems: a modelling approach. Ecological Modelling, 205, 492506.Google Scholar
Allen, A. P. and Gillooly, J. F. (2009). Towards an integration of ecological stoichiometry and the metabolic theory of ecology to better understand nutrient cycling. Ecology Letters, 12, 369384.Google Scholar
Allison, S. D. (2006). Brown ground: a soil carbon analogue for the green world hypothesis? American Naturalist, 167, 619627.Google Scholar
Beare, M. H., Parmelee, R. W., Hendrix, P. F., et al. (1992). Microbial and faunal interactions and effects on litter nitrogen and decomposition in agroecosystems. Ecological Monographs, 62, 569591.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
Bohan, D. A., Raybould, A., Mulder, C., et al. (2013). Networking agroecology: integrating the diversity of agroecosystem interactions. Advances in Ecological Research, 49, 167.Google Scholar
Cohen, J. E. (1978). Food Webs and Niche Space. Princeton, NJ: Princeton University Press.Google Scholar
Cohen, J. E. and Carpenter, S. R. (2005). Species’ average body mass and numerical abundance in a community food web: statistical questions in estimating the relationship. In Dynamic Food Webs: Multispecies Assemblages, Ecosystem Development, and Environmental Change, ed. de Ruiter, P. C., Wolters, V., and Moore, J. C., San Diego: Academic Press, pp. 137156.Google Scholar
Cohen, J. E. and Mulder, C. (2014). Soil invertebrates, chemistry, weather, human management, and edaphic food webs at 135 sites in the Netherlands: SIZEWEB. Ecology, 95, 578.Google Scholar
Coleman, D. C., Crossley, D. A. Jr., and Hendrix, P. F. (2004). Fundamentals of Soil Ecology, 2nd edn. San Diego: Academic Press.Google Scholar
De Visser, S. N., Freymann, B. P., and Olff, H. (2011). The Serengeti food web: empirical quantification and analysis of topological changes under increasing human impact. Journal of Animal Ecology, 80, 484494.Google Scholar
Duncan, C., Thompson, J. R., and Pettorelli, N. (2015). The quest for a mechanistic understanding of biodiversity–ecosystem services relationships. Proceedings of the Royal Society B: Biological Sciences, 282, 20151348.Google Scholar
Fitter, A. H., Gilligan, C. A., Hollingworth, K., et al. (2005). Biodiversity and ecosystem function in soil. Functional Ecology, 19, 369377.Google Scholar
Gray, C., Baird, D. J., Baumgartner, S., et al. (2014). Ecological networks: the missing links in biomonitoring science. Journal of Applied Ecology, 51, 14441449.Google Scholar
Hagen, M., Kissling, W. D., Rasmussen, C., et al. (2012). Biodiversity, species interactions and ecological networks in a fragmented world. Advances in Ecological Research, 46, 89210.Google Scholar
Hairston, N. G., Smith, F. E., and Slobodkin, L. B. (1960). Community structure, population control, and competition. American Naturalist, 94, 421425.Google Scholar
Hieber, M. and Gessner, M. O. (2002). Contribution of stream detrivores, fungi, and bacteria to leaf breakdown based on biomass estimates. Ecology, 83, 10261038.Google Scholar
Hladyz, S., Ǻbjörnsson, K., Chauvet, E., et al. (2011). Stream ecosystem functioning in an agricultural landscape: the importance of terrestrial–aquatic linkages. Advances in Ecological Research, 44, 211276.Google Scholar
Hooper, D. U., Chapin, F. S. III, Ewel, J. J., et al. (2005). Effects of biodiversity on ecosystem functioning: a consensus of current knowledge and needs for future research. Ecological Monographs, 75, 335.Google Scholar
Hsu, S. C., Liu, S. C., Huang, Y.-T., et al. (2009). Long-range southeastward transport of Asian biosmoke pollution: signature detected by aerosol potassium in Northern Taiwan. Journal of Geophysical Research: Atmospheres, 114, D14301.Google Scholar
Hudson, L. N., Emerson, R., Jenkins, G. B., et al. (2013). Cheddar: analysis and visualisation of ecological communities in R. Methods in Ecology and Evolution, 4, 99104.Google Scholar
Hunt, H. W., Coleman, D. C., Ingham, E. R., et al. (1987). The detrital food web in a shortgrass prairie. Biology and Fertility of Soils, 3, 5768.Google Scholar
Kapo, K. E., Holmes, C. M., Dyer, S. D., De Zwart, D., and Posthuma, L. (2014). Developing a foundation for eco-epidemiological assessment of aquatic ecological status over large geographic regions utilizing existing data resources and models. Environmental Toxicology and Chemistry, 33, 16651677.Google Scholar
Kaspari, M. and Weiser, M. (2007). The size–grain hypothesis: do macroarthropods see a fractal world? Ecological Entomology, 32, 279282.Google Scholar
Kaspari, M. and Yanoviak, S. P. (2009). Biogeochemistry and the structure of tropical brown food webs. Ecology, 90, 33423351.Google Scholar
Kattge, J., Díaz, S., Lavorel, S., et al. (2011). TRY: a global database of plant traits. Global Change Biology, 17, 29052935.Google Scholar
Lavorel, S., McIntyre, S., Landsberg, J., and Forbes, T. D. A. (1997). Plant functional classifications: from general groups to specific groups based on response to disturbance. Trends in Ecology and Evolution, 12, 474478.Google Scholar
Lavorel, S., Storkey, J., Bardgett, R. D., et al. (2013). A novel framework for linking functional diversity of plants and other trophic levels for the quantification of ecosystem services. Journal of Vegetation Science, 22, 942948.Google Scholar
Leitch, A. R., Leitch, I. J., Trimmer, M., Guignard, M. S., and Woodward, G. (2014). Impact of genomic diversity in river ecosystems. Trends in Plant Science, 19, 361366.Google Scholar
Levins, R. (1974). The qualitative analysis of partially specified systems. Annals of the New York Academy of Sciences, 231, 123138.Google Scholar
Loreau, M. (2009). Linking biodiversity and ecosystems: towards a unifying ecological theory. Philosophical Transactions of the Royal Society B: Biological Sciences, 365, 4960.Google Scholar
Loreau, M., Downing, A. L., Emmerson, M. C., et al. (2002). A new look at the relationship between diversity and stability. In Biodiversity and Ecosystem Functioning. Synthesis and Perspectives, ed. Loreau, M., Naeem, S., and Inchausti, P., Oxford: Oxford University Press, pp. 7991.Google Scholar
Macfadyen, S., Gibson, R. H., Symondson, W. O. C., and Memmott, J. (2011). Landscape structure influences modularity patterns in farm food webs: consequences for pest control. Ecological Applications, 21, 516524.Google Scholar
Mahowald, N., Jickells, T. D., Baker, A. R., et al. (2008). Global distribution of atmospheric phosphorus sources, concentrations and deposition rates, and anthropogenic impacts. Global Biogeochemical Cycles, 22, GB4026.Google Scholar
Martin, L. J., Blossey, B., and Ellis, E. (2012). Mapping where ecologists work: biases in the global distribution of terrestrial ecological observations. Frontiers in Ecology and the Environment, 10, 195201.Google Scholar
McMahon, T. A., Halstead, N. T., Johnson, S., et al. (2012). Fungicide-induced declines of freshwater biodiversity modify ecosystem functions and services. Ecology Letters, 15, 714722.Google Scholar
Moore, J. C. and de Ruiter, P. C. (2012). Energetic Food Webs: An Analysis of Real and Model Ecosystems. Oxford: Oxford University Press.Google Scholar
Moore, J. C., Berlow, E. L., Coleman, D. C., et al. (2004). Detritus, trophic dynamics and biodiversity. Ecology Letters, 7, 584600.Google Scholar
Mulder, C. (2010). Soil fertility controls the size–specific distribution of eukaryotes. Annals of the New York Academy of Sciences, 1195, E7481.Google Scholar
Mulder, C. and Elser, J. J. (2009). Soil acidity, ecological stoichiometry and allometric scaling in grassland food webs. Global Change Biology, 15, 27302738.Google Scholar
Mulder, C., Boit, A., Mori, S., et al. (2012). Distributional (in)congruence of biodiversity–ecosystem functioning. Advances in Ecological Research, 46, 188.Google Scholar
Mulder, C., Ahrestani, F. S., Bahn, M., et al. (2013). Connecting the green and brown worlds: elemental factors and trait-driven predictability of ecological networks. Advances in Ecological Research, 49, 67173.Google Scholar
Mulder, C., Bennett, E. M., Bohan, D. A., et al. (2015). Ten years later: revisiting priorities for science and society a decade after the Millennium Ecosystem Assessment. Advances in Ecological Research, 53, 153.Google Scholar
Neutel, A.-M. and Thorne, M. A. S. (2014). Interaction strengths in balanced carbon cycles and the absence of a relation between ecosystem complexity and stability. Ecology Letters, 17, 651661.Google Scholar
Peñuelas, J., Poulter, B., Sardans, J., et al. (2013). Human-induced nitrogen–phosphorus imbalances alter natural and managed ecosystems across the globe. Nature Communications, 4, 2934.Google Scholar
Perrings, C., Naeem, S., Ahrestani, F. S., et al. (2011). Ecosystem services, targets, and indicators for the conservation and sustainable use of biodiversity. Frontiers in Ecology and the Environment, 9, 512520.Google Scholar
Pocock, M. J. O., Evans, D. M., and Memmott, J. (2012). The robustness and restoration of a network of ecological networks. Science, 335, 973977.Google Scholar
Posthuma, L., Bjørn, A., Zijp, M. C., et al. (2014). Beyond safe operating space: finding chemical footprinting feasible. Environmental Science and Technology, 38, 60576059.Google Scholar
QUINTESSENCE Consortium (2016). Networking our way to better ecosystem service provision. Trends in Ecology and Evolution, 31. 10.1016/j.tree.2015.12.003.Google Scholar
Rzanny, M., Kuu, A., and Voigt, W. (2013). Bottom–up and top–down forces structuring consumer communities in an experimental grassland. Oikos, 122, 967976.Google Scholar
Scherber, C., Eisenhauer, N., Weisser, W. W., et al. (2010). Bottom–up effects of plant diversity on multitrophic interactions in a biodiversity experiment. Nature, 468, 553556.Google Scholar
Schröter, D., Wolters, V., and de Ruiter, P. C. (2003). C and N mineralisation in the decomposer food webs of a European forest transect. Oikos, 102, 294308.Google Scholar
Sechi, V., Brussaard, L., De Goede, R. G. M., Rutgers, M., and Mulder, C. (2015). Choice of resolution by functional trait or taxonomy affects allometric scaling in soil food webs. American Naturalist, 185, 142149.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
Sterner, R. W. and Elser, J. J. (2002). Ecological Stoichiometry. Princeton: Princeton University Press.Google Scholar
Stewart, R. I. A., Dossena, M., Bohan, D. A., et al. (2013). Mesocosm experiments as a tool for ecological climate-change research. Advances in Ecological Research, 48, 71181.Google Scholar
Strong, D. R. and Frank, K. T. (2010). Human involvement in food webs. Annual Review of Environment and Resources, 35, 123.Google Scholar
Struebig, M. J., Kingston, T., Petit, E. J., et al. (2011). Parallel declines in species and genetic diversity in tropical forest fragments. Ecology Letters, 14, 582590.Google Scholar
Thébault, E. and Fontaine, C. (2010). Stability of ecological communities and the architecture of mutualistic and trophic networks. Science, 329, 853856.Google Scholar
Urabe, J., Naeem, S., Raubenheimer, D., and Elser, J. J. (2010). The evolution of biological stoichiometry under global change. Oikos, 119, 737740.Google Scholar
Von Carlowitz, H. C. (1713). Sylvicultura oeconomica, oder Haußwirthliche Nachricht und Naturmäßige Anweisung zur wilden Baum-Zucht. Leipzig: Johann Friedrich Braun.Google Scholar
Von Liebig, J. (1840). Die Organische Chemie in ihrer Anwendung auf Agricultur und Physiologie. Braunschweig: Vieweg.Google Scholar
Wall, D. H., Nielsen, U. N., and Six, J. (2015). Soil biodiversity and human health. Nature, 528, 6976.Google Scholar
Wardle, D. A. (2002). Communities and Ecosystems: Linking the Aboveground and Belowground Components. Princeton: Princeton University Press.Google Scholar
Wolters, V., Silver, W. L., Bignell, D. E., et al. (2000). Effects of global changes on above- and belowground biodiversity in terrestrial ecosystems: implications for ecosystem functioning. BioScience, 50, 10891098.Google Scholar
Woodward, G., Gessner, M. O., Giller, P. S., et al. (2012). Continental-scale effects of nutrient pollution on stream ecosystem functioning. Science, 336, 14381440.Google Scholar

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.Google Scholar
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.Google Scholar
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.Google Scholar
Brose, U. (2008). Complex food webs prevent competitive exclusion among producer species. Proceedings of the Royal Society B: Biological Sciences, 275, 25072514.Google Scholar
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.Google Scholar
Crooks, K. R. and Soulé, M. E. (1999). Mesopredator release and avifaunal extinctions in a fragmented system. Nature, 400, 563566.Google 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.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, 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.Google 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.Google 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.Google 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.Google Scholar
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.Google 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.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, 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.Google 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.Google 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.Google 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.Google 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

References

Abraham, W.-R. (2014). Applications and impacts of stable isotope probing for analysis of microbial interactions. Journal of Appled Microbiology and Biotechnology, 98, 48174828.Google Scholar
Agustí, N., Bourguet, D., Spataro, T., et al. (2005). Detection, identification and geographical distribution of European corn borer larval parasitoids using molecular markers. Molecular Ecology, 14, 32673274.Google Scholar
Alvarez, H. M. and Steinbüchel, A. (2002). Triacylglycerols in prokaryotic microorganisms. Journal of Applied Microbiology and Biotechnology, 60, 367376.Google Scholar
Andújar, C., Arribas, P., Ruzicka, F., et al. (2015). Phylogenetic community ecology of soil biodiversity using mitochondrial metagenomics. Molecular Ecology, 24(14), 36033617. doi:10.1111/mec.13195.Google Scholar
Baiser, B., Russell, G. J., and Lockwood, J. L. (2010). Connectance determines invasion success via trophic interactions in model food webs. Oikos, 119, 19701976.Google Scholar
Baker, R., Buckland, A., and Sheaves, M. (2014). Fish gut content analysis: robust measures of diet composition. Fish and Fisheries, 15, 170177.Google Scholar
Bardgett, R. D., Cook, R., Yeates, G. W., and Denton, C. S. (1999). The influence of nematodes on below-ground processes in grassland ecosystems. Plant Soil, 212, 2333.Google Scholar
Berg, M., de Ruiter, P. C., Didden, W. A. M., et al. (2001). Community food web, decomposition and nitrogen mineralisation in a stratified Scots pine forest soil. Oikos 94, 130142.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
Bongers, T. (1990). The maturity index: an ecological measure of environmental disturbance based on nematode species composition. Oecologica, 83, 1419.Google Scholar
Bonkowski, M. (2004). Protozoa and plant growth: the microbial loop in soil revisited. New Phytologist, 162, 617631.Google Scholar
Boschker, H. T. S., Nold, S. C., Wellsbury, P., et al. (1998). Direct linking of microbial populations to specific biogeochemical processes by 13C-labelling of biomarkers. Nature, 392, 801805.Google Scholar
Boyer, S., Yeates, G. W., Wratten, S. D., Holyoake, A., and Cruickshank, R. H. (2011). Molecular and morphological analyses of faeces to investigate the diet of earthworm predators: example of a carnivorous land snail endemic to New Zealand. Pedobiologia (Jena), 54, 153158.Google Scholar
Brose, U. (2010). Body-mass constraints on foraging behaviour determine population and food-web dynamics. Functional Ecology, 24, 2834.Google Scholar
Brose, U., Jonsson, T., Berlow, E. L., et al. (2006). Consumer-resource body-size relationships in natural food webs. Ecology, 87, 24112417.Google Scholar
Brose, U., Ehnes, R. B., Rall, B. C., et al. (2008). Foraging theory predicts predator–prey energy fluxes. Journal of Animal Ecology, 77, 10721078.Google Scholar
Buse, T., Ruess, L., and Filser, J. (2013). New trophic biomarkers for Collembola reared on algal diets. Pedobiologia (Jena), 56, 153159.Google Scholar
Buse, T., Ruess, L., and Filser, J. (2014). Collembola gut passage shapes microbial communities in faecal pellets but not viability of dietary algal cells. Chemoecology, 24, 7984.Google Scholar
Chamberlain, P. M., Bull, I. D., Black, H. I. J., Ineson, P., and Evershed, R. P. (2005). Fatty acid composition and change in Collembola fed differing diets: identification of trophic biomarkers. Soil Biology and Biochemistry, 37, 16081624.Google Scholar
Chen, J., Ferris, H., Scow, K. M., and Graham, K. J. (2001). Fatty acid composition and dynamics of selected fungal-feeding nematodes and fungi. Comparative Biochemistry and Physiology Part B: Biochemistry and Molecular Biology, 130, 135144.Google Scholar
Clare, E. L. (2014). Molecular detection of trophic interactions: emerging trends, distinct advantages, significant considerations and conservation applications. Evolutionary Applications, 7(9), 11441157. doi:10.1111/eva.12225.Google Scholar
Deagle, B. E. and Tollit, D. J. (2007). Quantitative analysis of prey DNA in pinniped faeces: potential to estimate diet composition? Conservation Genetics, 8, 743747.Google Scholar
Deagle, B. E., Kirkwood, R., Jarman, S. N. (2009). Analysis of Australian fur seal diet by pyrosequencing prey DNA in faeces. Molecular Ecology, 18, 20222038.Google Scholar
Deagle, B. E., Jarman, S. N., Coissac, E., Pompanon, F., and Taberlet, P. (2014). DNA metabarcoding and the cytochrome c oxidase subunit I marker. Biology Letters, 10, dx.doi.org/10.1098/rsbl.2014.0562.Google Scholar
de Ruiter, P. C., Moore, J. C., Zwart, K. B., et al. (1993). Simulation of nitrogen mineralization in the below-ground food webs of two winter wheat fields. Journal of Applied Ecology, 30, 95106.Google Scholar
de Ruiter, P. C., Wolters, V., and Moore, J. C. (2005). Dynamics Food Webs: Multispecies Assemblages, Ecosystem Development and Environmental Change. San Diego, CA: Academic Press.Google Scholar
Drigo, B., Pijl, A. S., Duyts, H., et al. (2010). Shifting carbon flow from roots into associated microbial communities in response to elevated atmospheric CO2. Proceeding of the National Academy of Sciences of the USA, 107, 1093810942.Google Scholar
Eggers, T. and Jones, T. H. (2000). You are what you eat… or are you? Trends in Ecology and Evolution, 15, 265266.Google Scholar
Ferlian, O., Scheu, S., and Pollierer, M. M. (2012). Trophic interactions in centipedes (Chilopoda, Myriapoda) as indicated by fatty acid patterns: variations with life stage, forest age and season. Soil Biology and Biochemistry, 52, 3342.Google Scholar
Ferris, H. (2010a). Contribution of nematodes to the structure and function of the soil food web. Journal of Nematology, 42, 6367.Google Scholar
Ferris, H. (2010b). Form and function: metabolic footprints of nematodes in the soil food web. European Journal of Soil Biology, 46, 97104.Google Scholar
Ferris, H., Bongers, T., and De Goede, R. G. M. (2001). A framework for soil food web diagnostics: extension of the nematode faunal analysis concept. Applied Soil Ecology, 18, 1329.Google Scholar
Freckman, D. W. (1988). Bacterivorous nematodes and organic-matter decomposition. Agriculture Ecosystems and Environment, 24, 195217.Google Scholar
Friedrich, M. W. (2011). Trophic interactions in microbial communities and food webs traced by stable isotope probing of nucleic acids. In Stable Isotope Probing and Related Technologies, ed. Murrell, J. C. and Whiteley, A. S., American Society for Microbiology Press, pp. 203232.Google Scholar
Gariepy, T., Kuhlmann, U., Gillott, C., and Erlandson, M. (2008). A large-scale comparison of conventional and molecular methods for the evaluation of host–parasitoid associations in non-target risk-assessment studies. Journal of Applied Ecology, 45, 708715.Google Scholar
Ghafouri, S. and McGhee, J. (2007). Bacterial residence time in the intestine of Caenorhabditis elegans. Nematology, 9, 8791.Google Scholar
Greenstone, M. H., Rowley, D. L., Weber, D. C., Payton, M. E., and Hawthorne, D. J. (2007). Feeding mode and prey detectability half-lives in molecular gut-content analysis: an example with two predators of the Colorado potato beetle. Bulletin of Entomological Research, 97, 201209.Google Scholar
Greenstone, M. H., Payton, M. E., Weber, D. C., and Simmons, A. M. (2014). The detectability half-life in arthropod predator–prey research: what it is, why we need it, how to measure it, and how to use it. Molecular Ecology, 23, 37993813.Google Scholar
Griffiths, B. S., Ritz, K., Ebblewhite, N., and Dobson, G. (1999). Soil microbial community structure: effects of substrate loading rates. Soil Biology and Biochemistry, 31, 145153.Google Scholar
Harper, G. L., King, R. A., Dodd, C. S., et al. (2005). Rapid screening of invertebrate predators for multiple prey DNA targets. Molecular Ecology, 14, 819827.Google Scholar
Haubert, D., Häggblom, M. M., Scheu, S., and Ruess, L. (2004). Effects of fungal food quality and starvation on the fatty acid composition of Protaphorura fimata (Collembola). Comparative Biochemistry and Physiology Part B: Biochemistry and Molecular Biology, 138, 4152.Google Scholar
Haubert, D., Häggblom, M. M., Langel, R., Scheu, S., and Ruess, L. (2006). Trophic shift of stable isotopes and fatty acids in Collembola on bacterial diets. Soil Biology and Biochemistry, 38, 20042007.Google Scholar
Haubert, D., Häggblom, M. M., Scheu, S., and Ruess, L. (2008). Effects of temperature and life stage on the fatty acid composition of Collembola. European Journal of Soil Biology, 44, 213219.Google Scholar
Haubert, D., Birkhofer, K., Fließbach, A., et al. (2009). Trophic structure and major trophic links in conventional versus organic farming systems as indicated by carbon stable isotope ratios of fatty acids. Oikos, 118, 15791589.Google Scholar
Heidemann, K., Scheu, S., Ruess, L., and Maraun, M. (2011). Molecular detection of nematode predation and scavenging in oribatid mites: laboratory and field experiments. Soil Biology and Biochemistry, 43, 22292236.Google Scholar
Heidemann, K., Hennies, A., Schakowske, J., et al. (2014a). Free-living nematodes as prey for higher trophic levels of forest soil food webs. Oikos, 123, 11991211.Google Scholar
Heidemann, K., Ruess, L., Scheu, S., and Maraun, M. (2014b). Nematode consumption by mite communities varies in different forest microhabitats as indicated by molecular gut content analysis. Experimental and Applied Acarology, 64, 4960.Google Scholar
Holmstrup, M., Hedlund, K., and Boriss, H. (2002). Drought acclimation and lipid composition in Folsomia candida: implications for cold shock, heat shock and acute desiccation stress. Journal of Insect Physiology, 48, 961970.Google Scholar
Hrček, J. and Godfray, H. C. J. (2015). What do molecular methods bring to host–parasitoid food webs? Trends in Parasitology, 31, 3035.Google Scholar
Huang, W. E., Stoecker, K., Griffiths, R., et al. (2007). Raman–FISH: combining stable-isotope Raman spectroscopy and fluorescence in situ hybridization for the single cell analysis of identity and function. Environmental Microbiology, 9, 18781889.Google Scholar
Hunt, H. W., Coleman, D. C., Ingham, E. R., et al. (1987). The detrital food web in a shortgrass prairie. Biology and Fertility of Soils, 3, 5768.Google Scholar
Hyslop, E. J. (1980). Stomach contents analysis: a review of methods and their application. Journal of Fish Biology, 17, 411429.Google Scholar
Iverson, S. J., Field, C., Bowen, W. D., and Blanchard, W. (2004). Quantitative fatty acid signature analysis: a new method of estimating predator diets. Ecological Monographs, 74, 211235.Google Scholar
Jezbera, J., Hornák, K., and Simek, K. (2005). Food selection by bacterivorous protists: insight from the analysis of the food vacuole content by means of fluorescence in situ hybridization. FEMS Microbiol Ecology, 52, 351363.Google Scholar
Jonsson, T. (2014). Trophic links and the relationship between predator and prey body sizes in food webs. Community Ecology, 15, 5464.Google Scholar
Jousset, A. (2012). Ecological and evolutive implications of bacterial defences against predators. Environmental Microbiology, 14, 18301843.Google Scholar
Jousset, A., Rochat, L., Péchy-Tarr, M., et al. (2009). Predators promote defence of rhizosphere bacterial populations by selective feeding on non-toxic cheaters. ISME Journal, 3, 666674.Google Scholar
Juen, A. and Traugott, M. (2005). Detecting predation and scavenging by DNA gut-content analysis: a case study using a soil insect predator–prey system. Oecologia, 142, 344352.Google Scholar
Juen, A. and Traugott, M. (2007). Revealing species-specific trophic links in soil food webs: molecular identification of scarab predators. Molecular Ecology, 16, 15451557.Google Scholar
Jürgens, K. and Simek, K. (2000). Functional response and particle size selection of Halteria cf. grandinella, a common freshwater oligotrichous ciliate. Aquatic Microbial Ecology, 22, 5768.Google Scholar
Kalinkat, G., Rall, B. C., Vucic-Pestic, O., and Brose, U. (2011). The allometry of prey preferences. PLoS One, 6, e25937.Google Scholar
Kasper, M. L., Reeson, A. F., Cooper, S. J. B, Perry, K. D., and Austin, A. D. (2004). Assessment of prey overlap between a native (Polistes humilis) and an introduced (Vespula germanica) social wasp using morphology and phylogenetic analyses of 16S rDNA. Molecular Ecology, 13, 20372048.Google Scholar
King, R. A., Read, D. S., Traugott, M., and Symondson, W. O. C. (2008). Molecular analysis of predation: a review of best practice for DNA-based approaches. Molecular Ecology, 17, 947963.Google Scholar
King, R. A., Vaughan, I. P., Bell, J. R., Bohan, D. A., and Symondson, W. O. C. (2010). Prey choice by carabid beetles feeding on an earthworm community analysed using species- and lineage-specific PCR primers. Molecular Ecology, 19, 17211732.Google Scholar
Layman, C. A., Araujo, M. S., Boucek, R., et al. (2012). Applying stable isotopes to examine food-web structure: an overview of analytical tools. Biological Reviews, 87, 545562.Google Scholar
Li, M., Huang, W. E., Gibson, C. M., Fowler, P. W., and Jousset, A. (2013). Stable isotope probing and Raman spectroscopy for monitoring carbon flow in a food chain and revealing metabolic pathway. Analytical Chemistry, 85, 16421649.Google Scholar
Lundgren, J. G. and Fergen, J. K. (2014). Predator community structure and trophic linkage strength to a focal prey. Molecular Ecology, 23, 37903798.Google Scholar
MacArthur, R. H. (1972). Strong, or weak, interations? Transactions of the Connecticut Academy of Arts and Sciences, 44, 177188.Google Scholar
Manefield, M., Whiteley, A. S., Griffiths, R. I., and Bailey, M. J. (2002). RNA stable isotope probing, a novel means of linking microbial community function to phylogeny. Rapid Communication in Mass Spectrometry, 16, 21792183.Google Scholar
Maxfield, P. J. and Evershed, R. P. (2011). Phospholipid fatty acid stable isotope probing techniques in microbial ecology. In Stable Isotope Probing and Related Technologies, ed. Murrell, J. C. and Whiteley, A. S., American Society of Microbiology Press, pp. 3771.Google Scholar
May, R. M. (1972). Will a large complex system be stable? Nature, 238, 413414.Google Scholar
May, R. M. (1974). Stability and Complexity in Model Ecosystems. Princeton, NJ: Princeton University Press.Google Scholar
Mayali, X., Weber, P. K., Brodie, E. L., et al. (2012). High-throughput isotopic analysis of RNA microarrays to quantify microbial resource use. ISME Journal, 6, 12101221.Google Scholar
Memmott, J. (2009). Food webs: a ladder for picking strawberries or a practical tool for practical problems? Philosophical Transactions of the Royal Society B: Biological Sciences, 364, 16931699.Google Scholar
Molkentin, J. and Giesemann, A. (2007). Differentiation of organically and conventionally produced milk by stable isotope and fatty acid analysis. Analytical Bioanalytical Chemistry, 388, 297305.Google Scholar
Moore, J. C. and de Ruiter, P. C. (2012). Energetic Food Webs. Oxford, UK: Oxford University Press.Google Scholar
Moore, J. C., de Ruiter, P. C., and Hunt, H. W. (1993). Soil invertebrate/micro-invertebrate interactions: disproportionate effects of species on food web structure and function. Veterinary Parasitology, 48, 247260.Google Scholar
Mora, C. A., Paunescu, D., Grass, R. N., and Stark, W. J. (2014). Silica particles with encapsulated DNA as trophic tracers. Molecular Ecology Resources, 15(2), 231241, doi:10.1111/1755–0998.12299.Google Scholar
Müller-Navarra, D. C., Brett, M. T., Liston, A. M., and Goldman, C. R. (2000). A highly unsaturated fatty acid predicts carbon transfer between primary producers and consumers. Nature, 403, 7477.Google Scholar
Neher, D. A. (2010). Ecology of plant and free-living nematodes in natural and agricultural soil. Annual Review of Phytopathology, 48, 371394.Google Scholar
Neidig, N., Paul, R. J., Scheu, S., and Jousset, A. (2011). Secondary metabolites of Pseudomonas fluorescens CHA0 drive complex non-trophic interactions with bacterivorous nematodes. Microbial Ecology, 61, 853859.Google Scholar
Neutel, A.-M., Heesterbeek, J. A. P., and de Ruiter, P. C. (2002). Stability in real food webs: weak links in long loops. Science, 296, 11201123.Google Scholar
Ngosong, C., Gabriel, E., and Ruess, L. (2012). Use of the signature fatty acid 16:1ω5 as a tool to determine the distribution of arbuscular mycorrhizal fungi in soil. Journal of Lipids, 2012. doi:10.1155/2012/236807.Google Scholar
Ngosong, C., Gabriel, E., and Ruess, L. (2014). Collembola grazing on arbuscular mycorrhiza fungi modulates nutrient allocation in plants. Pedobiologia (Jena), 57, 171179.Google Scholar
O’Neill, R. V. (1969). Indirect estimation of energy fluxes in animal food webs. Journal of Theoretical Biology, 22, 284290.Google Scholar
Paine, R. T. (1980). Food webs: linkage, interation strength and community infrastructure. Journal of Animal Ecology, 49, 667685.Google Scholar
Paine, R. T. (1992). Food-web analysis through field measurement of per capita interaction strength. Nature, 355, 7375.Google Scholar
Pausch, J., Kramer, S., Scharroba, A., et al. (2015). Small but active: pool size does not matter for carbon incorporation in belowground food webs. Functional Ecology, 30(3), 479489.Google Scholar
Pimm, S. L. (1982). Food Webs. London, UK: Chapman and Hall.Google Scholar
Pollierer, M. M., Scheu, S., and Haubert, D. (2010). Taking it to the next level: trophic transfer of marker fatty acids from basal resource to predators. Soil Biology and Biochemistry, 42, 919925.Google Scholar
Pompanon, F., Deagle, B. E., Symondson, W. O. C., et al. (2012). Who is eating what: diet assessment using next generation sequencing. Molecular Ecology, 21, 19311950.Google Scholar
Pond, D. W., Leakey, R. J. G., and Fallick, A. E. (2006). Monitoring microbial predator–prey interactions: an experimental study using fatty acid biomarker and compound-specific stable isotope techniques. Journal of Plankton Research, 28, 419427.Google Scholar
Radajewski, S., Ineson, P., Parekh, N. R., and Murrell, J. C. (2000). Stable-isotope probing as a tool in microbial ecology. Nature, 403, 646649.Google Scholar
Ruess, L. (2003). Nematode soil faunal analysis of decomposition pathways in different ecosystems. Nematology, 5, 179181.Google Scholar
Ruess, L. and Chamberlain, P. M. (2010). The fat that matters: soil food web analysis using fatty acids and their carbon stable isotope signature. Soil Biology and Biochemistry, 42, 18981910.Google Scholar
Ruess, L. and Ferris, H. (2004). Decomposition pathways and successional changes. Nematology Monographs and Perspectives, 2, 547556.Google Scholar
Ruess, L., Häggblom, M. M., Garciá Zapata, E. J., and Dighton, J. (2002). Fatty acids of fungi and nematodes: possible biomarkers in the soil food chain? Soil Biology and Biochemistry, 34, 745756.Google Scholar
Ruess, L., Häggblom, M. M., Langel, R., and Scheu, S. (2004). Nitrogen isotope ratios and fatty acid composition as indicators of animal diets in belowground systems. Oecologia, 139, 336346.Google Scholar
Ruess, L., Tiunov, A., Haubert, D., et al. (2005a). Carbon stable isotope fractionation and trophic transfer of fatty acids in fungal based soil food chains. Soil Biology Biochemistry, 37, 945953.Google Scholar
Ruess, L., Schütz, K., Haubert, D., et al. (2005b). Application of lipid analysis to understand trophic interactions in soil. Ecology, 86, 20752082.Google Scholar
Ruess, L., Schütz, K., Migge-Kleian, S., et al. (2007). Lipid composition of Collembola and their food resources in deciduous forest stands: implications for feeding strategies. Soil Biology and Biochemistry, 39, 19902000.Google Scholar
Sanders, D., Jones, C. G., Thébault, E., et al. (2014). Integrating ecosystem engineering and food webs. Oikos, 123, 513524.Google Scholar
Scheu, S. (2002). The soil food web: structure and perspectives. European Journal of Soil Biology, 38, 1120.Google Scholar
Scheu, S., Ruess, L., and Bonkowski, M. (2005). Interactions between microorganisms and soil micro- and mesofauna. In Microorganisms in Soil: Roles in Genesis and Functions, ed. Buscot, F. and Varma, A., Berlin: Springer-Verlag, pp. 253277.Google Scholar
Schröter, D., Wolters, V., and de Ruiter, P. C. (2003). C and N mineralisation in the decomposer food webs of a European forest transect. Oikos, 102, 294308.Google Scholar
Seeber, J., Rief, A., Seeber, G. U. H., Meyer, E., and Traugott, M. (2010). Molecular identification of detritivorous soil invertebrates from their faecal pellets. Soil Biology Biochemistry, 42, 12631267.Google Scholar
Selakovic, S., de Ruiter, P. C., and Heesterbeek, H. (2014). Infectious disease agents mediate interaction in food webs and ecosystems. Proceedings of the Royal Society B: Biological Sciences, 281(1777), 20132709.Google Scholar
Sint, D., Raso, L., and Traugott, M. (2012). Advances in multiplex PCR: balancing primer efficiencies and improving detection success. Methods in Ecology and Evolution, 3, 898905.Google Scholar
Stübing, D., Hagen, W., and Schmidt, K. (2003). On the use of lipid biomarkers in marine food web analyses: an experimental case study on the Antarctic krill, Euphausia superba. Limnology and Oceanography, 48, 16851700.Google Scholar
Symondson, W. O. C. (2012). The molecular revolution: using polymerase chain reaction based methods to explore the role of predators in terrestrial food webs. In Biodiversity and Insect Pests: Key Issues for Sustainable Management, ed. Gurr, G. M., Wratten, S. D., Snyder, W. E., and Read, D. M. Y., New York: Wiley and Sons, pp. 166184.Google Scholar
Thiemann, G. W., Iverson, S. J., and Stirling, I. (2008). Polar bear diets and arctic marine food webs: insights from fatty acid analysis. Ecological Monographs, 78, 591613.Google Scholar
Traugott, M., Kamenova, S., and Ruess, L. (2013). Empirically characterising trophic networks: what emerging DNA-based methods, stable isotope and fatty acid analyses can offer. Advances in Ecological Research, 49, 177224, doi:10.1016/B978-0–12-420002–9.00003–2.Google Scholar
van Dooremalen, C. and Ellers, J. (2010). A moderate change in temperature induces changes in fatty acid composition of storage and membrane lipids in a soil arthropod. Journal Insect Physiology, 56, 178184.Google Scholar
Vestheim, H. and Jarman, S. N. (2008). Blocking primers to enhance PCR amplification of rare sequences in mixed samples: a case study on prey DNA in Antarctic krill stomachs. Frontiers in Zoology, 5, 12, DOI: 10.1186/1742–9994-5–12.Google Scholar
von Berg, K., Traugott, M., and Scheu, S. (2012). Scavenging and active predation in generalist predators: a mesocosm study employing DNA-based gut content analysis. Pedobiologia (Jena), 55, 15.Google Scholar
Vuvic-Pestic, O. K., Birkhofer, K., Rall, B. C., Scheu, S., and Brose, U. (2010). Habitat structure and prey aggregation determine the functional response in a soil predator–prey interaction. Pedobiologica, 53, 307312.Google Scholar
Waldner, T. and Traugott, M. (2012). DNA-based analysis of regurgitates: a noninvasive approach to examine the diet of invertebrate consumers. Molecular Ecology Resoures, 12, 669675.Google Scholar
Waldner, T., Sint, D., Juen, A., and Traugott, M. (2013). The effect of predator identity on post-feeding prey DNA detection success in soil-dwelling macro-invertebrates. Soil Biology Biochemistry, 63, 116123.Google Scholar
Wallinger, C., Staudacher, K., Schallhart, N., et al. (2013). The effect of plant identity and the level of plant decay on molecular gut content analysis in a herbivorous soil insect. Molecular Ecology Resoures, 13, 7583.Google Scholar
Wardle, D. A., Verhoef, H. A., and Clarholm, M. (1998). Trophic relationships in the soil microfood-web: predicting the responses to a changing global environment. Global Change Biology, 4, 713727.Google Scholar
White, D. C., Stair, J. O., and Ringelberg, D. B. (1996). Quantitative comparisons of in situ microbial biodiversity by signature biomarker analysis. Journal of Industrial Microbiology, 17, 185196.Google Scholar
Williams, C. T. and Buck, C. L. (2010). Using fatty acids as dietary tracers in seabird trophic ecology: theory, application and limitations. Journal of Ornithology, 151, 531543.Google Scholar
Winemiller, K. O. and Layman, C. A. (2005). Food web science: moving on the path from abstraction to prediction. In Dynamic Food Webs, ed. de Ruiter, P. C., Moore, J. C., and Wolters, V., San Diego: Academic Press, pp. 1023.Google Scholar
Wirta, H. K., Hebert, P. D. N., Kaartinen, R., et al. (2014). Complementary molecular information changes our perception of food web structure. Proceedings of the National Academy of Sciences of the United States of America, 111, 18851890.Google Scholar
Yeates, G. W. (2010). Nematodes in ecological webs. In Encyclopedia of Life Sciences (ELS), New York: John Wiley and Sons. doi:10.1002/9780470015902.a0021913.Google Scholar
Yeates, G. W., Bongers, T., De Goede, R. G., Freckman, D. W., and Georgieva, S. S. (1993). Feeding habits in soil nematode families and genera: an outline for soil ecologists. Journal of Nematology, 25, 315331.Google Scholar
Yeates, G. W., Ferris, H., Moens, T., and van der Putten, W. H. (2009). The role of nematodes in ecosystems. In Nematodes as Environmental Bioindicators, ed. Wilson, M. and Kakouli-Duarte, T., Wallingford, UK: CABI, pp. 144.Google Scholar
Zelles, L. (1999). Fatty acid patterns of phospholipids and lipopolysaccharides in the characterisation of microbial communities in soil: a review. Biology and Fertility of Soils, 29, 111129.Google Scholar
Zhang, G.-F., , Z.-C., Wan, F.-H., and Lövei, G. L. (2007). Real-time PCR quantification of Bemisia tabaci (Homoptera: Aleyrodidae) B-biotype remains in predator guts. Molecular Ecology Notes, 7, 947954.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
×