Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-25T20:39:01.970Z Has data issue: false hasContentIssue false

10 - Importance of Trait-Related Flexibility for Food-Web Dynamics and the Maintenance of Biodiversity

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. 146 - 163
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. 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.CrossRefGoogle ScholarPubMed
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.CrossRefGoogle ScholarPubMed
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.Google Scholar
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.CrossRefGoogle 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

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
×