Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-2xdlg Total loading time: 0 Render date: 2024-06-19T04:54:06.926Z Has data issue: false hasContentIssue false

9 - Including the Life Cycle in Food Webs

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

Published online by Cambridge University Press:  05 December 2017

John C. Moore
Affiliation:
Colorado State University
Peter C. de Ruiter
Affiliation:
Wageningen Universiteit, The Netherlands
Kevin S. McCann
Affiliation:
University of Guelph, Ontario
Volkmar Wolters
Affiliation:
Justus-Liebig-Universität Giessen, Germany
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Adaptive Food Webs
Stability and Transitions of Real and Model Ecosystems
, pp. 121 - 145
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

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.CrossRefGoogle ScholarPubMed
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.Google Scholar
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.CrossRefGoogle ScholarPubMed
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.CrossRefGoogle 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.Google 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.CrossRefGoogle 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.CrossRefGoogle ScholarPubMed
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.Google Scholar
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.CrossRefGoogle 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.Google 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.CrossRefGoogle 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.Google 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.Google Scholar
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.Google Scholar
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.Google 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.CrossRefGoogle 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.Google Scholar
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.CrossRefGoogle ScholarPubMed
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

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
×