Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-xm8r8 Total loading time: 0 Render date: 2024-06-22T11:02:12.006Z Has data issue: false hasContentIssue false

18 - Food-Web Dynamics When Divergent Life-History Strategies Respond to Environmental Variation Differently: A Fisheries Ecology Perspective

from Part III - Food Webs and Environmental Sustainability

Published online by Cambridge University Press:  05 December 2017

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

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Adaptive Food Webs
Stability and Transitions of Real and Model Ecosystems
, pp. 305 - 323
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

Alcántara, J. M. and Rey, P. J. (2012). Linking topological structure and dynamics in ecological networks. American Naturalist, 180, 186199.Google Scholar
Allain, V. (2005). Ecopath model of the pelagic ecosystem of the Western and Central Pacific Ocean. First Regular Session of the Scientific Committee of the Western and Central Pacific Fisheries Commission, WCPFC-SC1-EB WP-10, 119.Google Scholar
Allain, V., Nicol, S., Essington, T., et al. (2007). An Ecopath with Ecosim model of the Western and Central Pacific Ocean warm pool pelagic ecosystem. Third Regular Session of the Scientific Committee of the Western and Central Pacific Fisheries Commission, WCPFC-SC3-EB SWG/IP-8, 142.Google Scholar
Allesina, S. and Pascual, M. (2008). Network structure, predator–prey modules, and stability in large food webs. Theoretical Ecology, 1, 5564.Google Scholar
Andrewartha, H. G. and Birch, C. (1954). The Distribution and Abundance of Animals. Chicago, IL: University of Chicago Press.Google Scholar
Angert, A. L., Huxman, T. E., Chesson, P., and Venable, D. L. (2009). Functional tradeoffs determine species coexistence via the storage effect. Proceedings of the National Academy of Sciences of the United States of America, 106, 1164111645.Google Scholar
Atkinson, A., Hill, S. L., Barange, M., et al. (2014). Sardine cycles, krill declines, and locust plagues: revisiting “wasp-waist” food webs. Trends in Ecology and Evolution, 29, 309316.Google Scholar
Bailly, D., Agostinho, A. A., and Suzuki, H. I. (2008). Influence of the flood regime on the reproduction of fish species with different reproductive strategies in the Cuiabá River, Upper Pantanal, Brazil. River Research and Applications, 24, 12181229.Google Scholar
Bakun, A. and Broad, K. (2003). Environmental “loopholes” and fish population dynamics: comparative pattern recognition with focus on El Niño effects in the Pacific. Fisheries Oceanography, 12, 458473.Google Scholar
Barton, B. T. and Ives, A. R. (2014). Species interactions and a chain of indirect effects driven by reduced precipitation. Ecology, 95, 486494.Google Scholar
Bellmore, J. R., Baxter, C. V., and Connolly, P. J. (2015). Spatial complexity reduces interaction strengths in the meta-food web of a river floodplain mosaic. Ecology, 96, 274283.Google Scholar
Camerano, L. (1880). Dell’equilibrio dei viventi merce la reciproca distribuzione. Atti della Scienze di Torino, 15, 393414.Google Scholar
Carpenter, S. R. and Turner, M. G. (2001). Hares and tortoises: interactions of fast and slow variables in ecosystems. Ecosystems, 3, 495497.Google Scholar
Cheal, A. J., Delean, S., and Thompson, A. A. (2007). Spatial synchrony in coral reef fish populations and the influence of climate. Ecology, 88, 158169.Google Scholar
Chesson, P. (2000). General theory of competitive coexistence in spatially varying environments. Theoretical Population Biology, 58, 211237.CrossRefGoogle ScholarPubMed
Chesson, P. and Warner, R. (1981). Environmental variability promotes coexistence in lottery competitive systems. American Naturalist, 117, 923943.Google Scholar
Christensen, V. (2013). Ecological networks in fisheries: predicting the future? Fisheries, 38, 7681.Google Scholar
Christensen, V. and Walters, C. J. (2004). Ecopath with Ecosim: methods, capabilities and limitations. Ecological Modelling, 172(2–4), 109139.Google Scholar
De Roos, A. M., Persson, L., and McCauley, E. (2003). The influence of size-dependent life history traits on the structure and dynamics of populations and communities. Ecology Letters, 6, 473487.Google Scholar
Fujiwara, M. (2016). Incorporating demographic diversity into food web models: effects on community structure and dynamics. Ecological Modelling, 322, 1018.Google Scholar
Giacomini, H. C., DeAngelis, D. L., Trexler, J. C., and Petrere, M. Jr (2013). Trait contributions to fish community assembly emerge from trophic interactions in an individual-based model. Ecological Modelling, 251, 3243.Google Scholar
Grime, J. P. (1977). Evidence for existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. American Naturalist, 111, 11691194.Google Scholar
Grimm, V., Revilla, E., Berger, U., et al. (2005). Pattern-oriented modeling of agent-based complex systems: lessons from ecology. Science, 310, 987991.Google Scholar
Hartvig, M., Andersen, K. H., and Beyer, J. E. (2011). Food web framework for size-structured populations. Journal of Theoretical Biology, 272, 113122.Google Scholar
Hastings, A. (1996). What equilibrium behavior of Lotka–Volterra models does not tell us about food webs. In Food Webs: Integration of Patterns and Dynamics, ed. Polis, G. A. and Winemiller, K. O., New York, NY: Chapman & Hall, pp. 211217.Google Scholar
Hastings, A., Hom, C., Ellner, S., Turchin, P., and Godfray, H. C. J. (1993). Chaos in ecology: is mother nature a strange attractor? Annual Reviews of Ecology and Systematics, 24, 133.Google Scholar
Haydon, D. (1994). Pivotal assumptions determining the relationship between stability and complexity: an analytical synthesis of the stability-complexity debate. American Naturalist, 144, 1429.Google Scholar
Higgins, S., Pickett, S. T. A., and Bond, W. J. (2000). Predicting extinction risks for plants: environmental stochasticity can save declining populations. Trends in Ecology and Evolution, 15, 516520.Google Scholar
Holt, R. D. (2008). Theoretical perspectives on resource pulses. Ecology, 89, 671681.Google Scholar
Holt, R. D. (2009). Towards a trophic island biogeography: reflections on the interface of island biogeography and food web ecology. In The Theory of Island Biogeography Revisited, ed. Losos, J. B. and Ricklefs, R. E., Princeton, NJ: Princeton University Press, pp. 143185.Google Scholar
Jacobsen, N. S., Gislason, H., and Andersen, K. H. (2014). The consequences of balanced harvesting of fish communities. Proceedings of the Royal Society B: Biological Sciences, 281(1775), DOI: 10.1098/rspb.2013.2701.Google Scholar
Koenig, W. D. and Liebhold, A. M. (2005). Effects of periodical cicada emergences on abundance and synchrony of avian populations. Ecology, 86, 18731882.Google Scholar
Lawton, J. H. (1995). Population dynamic principles. In Extinction Rates, ed. Lawton, J. H. and May, R. M., Oxford, UK: Oxford University Press, pp. 147163.Google Scholar
Lindeman, R. L. (1942). The trophic-dynamic aspect of ecology. Ecology, 23, 399418.CrossRefGoogle Scholar
Madsen, T. and Shine, R. (2000). Rain, fish and snakes: climatically driven population dynamics of Arafura filesnakes in tropical Australia. Oecologia, 124, 208215.Google Scholar
May, R. M. (1973). Stability and Complexity in Model Ecosystems. Princeton, NJ: Princeton University Press.Google Scholar
McCann, K. (2012). Food Webs. Princeton, NJ: Princeton University Press.Google Scholar
Melián, C. J., Baldó, F., Matthews, B., et al. (2014). Individual trait variation and diversity in food webs. Advances in Ecological Research, 50, 207241.Google Scholar
Menge, B. A., Gouhier, T., Friedenburg, T., et al. (2011). Linking long-term, large-scale climatic and environmental variability to patterns of marine invertebrate recruitment: toward explaining “unexplained” variation. Journal of Experimental Marine Biology and Ecology, 400, 236249.Google Scholar
Olff, H., Alonso, D., Berg, M. P., et al. (2009). Parallel ecological networks in ecosystems. Philosophical Transactions of the Royal Society B: Biological Sciences, 364, 17551779.Google Scholar
Paine, R. T. (1988). On food webs: road maps of interactions or the grist for theoretical development? Ecology, 69, 16481654.Google Scholar
Pianka, E. R. (1970). On r- and K-selection. American Naturalist, 104, 592597.Google Scholar
Polis, G. A., Holt, R. D., Menge, B. A., and Winemiller, K. O. (1996). Time, space, and life history: influences on food webs. In Food Webs: Integration of Patterns and Dynamics, ed. Polis, G. A. and Winemiller, K. O., New York, NY: Chapman and Hall, pp. 435460.Google Scholar
Post, D. M. (2002). The long and short of food-chain length. Trends in Ecology and Evolution, 17, 269277.Google Scholar
Power, M. E., Parker, M. S., and Dietrich, W. E. (2008). Seasonal reassembly of a river food web: floods, droughts, and impacts of fish. Ecological Monographs, 78, 263282.Google Scholar
Rose, K. A. (2012). End-to-end models for marine ecosystems: are we on the precipice of a significant advance or just putting lipstick on a pig? Scientia Marina, 76, 195201.Google Scholar
Rose, K. A., Cowan, J. H., Winemiller, K. O., Myers, R. A., and Hilborn, R. (2001). Compensatory density-dependence in fish populations: importance, controversy, understanding, and prognosis. Fish and Fisheries, 2, 293327.Google Scholar
Rudolf, V. H. W. and Lafferty, K. D. (2011). Stage structure alters how complexity affects stability of ecological networks. Ecology Letters, 14, 7579.Google Scholar
Schmitz, O. J. and Booth, G. (1997). Modelling food web complexity: the consequences of individual-based, spatially explicit behavioural ecology on trophic interactions. Evolutionary Ecology, 11, 379398.Google Scholar
Siddon, E. C., Kristiansen, T., Mueter, F. J., et al. (2013). Spatial match-mismatch between juvenile fish and prey provides a mechanism for recruitment variability across contrasting climate conditions in the Eastern Bering Sea. PLOS One, 8(12), e84526.Google Scholar
Sinclair, A. R. E. (2003). Mammal population regulation, keystone processes and ecosystem dynamics. Philosophical Transactions of the Royal Society B: Biological Sciences, 358, 17291740.Google Scholar
Soberón, J. (2007). Grinnellian and Eltonian niches and geographic distributions of species. Ecology Letters, 10, 11151123.Google Scholar
Southwood, T. R. E. (1977). Habitat, the templet for ecological strategies? Journal of Animal Ecology, 46, 337365.Google Scholar
Stearns, S. C. (1992). The Evolution of Life Histories. Oxford, UK: Oxford University Press.Google Scholar
Summerhayes, V. S. and Elton, C. S. (1923). Contribution to the ecology of Spitsbergen and Bear Island. Journal of Ecology, 11, 214286.CrossRefGoogle Scholar
Tang, S., Pawar, S., and Allesina, S. (2014). Correlation between interaction strengths drives stability in large ecological networks. Ecology Letters, 17, 10941100.Google Scholar
Travers, M., Shin, Y.-J., Jennings, S., and Cury, P. (2007). Towards end-to-end models for investigating the effects of climate and fishing in marine ecosystems. Progressive Oceanography, 75, 751770.Google Scholar
Travis, J., Coleman, F. C., Auster, P. J., et al. (2014). Integrating the invisible fabric of nature into fisheries management. Proceedings of the National Academy of Sciences of the United States of America, 111(2), 581584.Google Scholar
Turchin, P. (2003). Complex Population Dynamics. A Theoretical/Empirical Synthesis. Princeton, NJ: Princeton University Press.Google Scholar
van Kooten, T., de Roos, A. M., and Persson, L. (2005). Bistability and an Allee effect as emergent consequences of stage-specific predation. Journal of Theoretical Biology, 237, 6774.Google Scholar
Van Winkle, W., Rose, K. A., and Chambers, R. C. (1993). Individual-based approach to fish population dynamics: an overview. Transactions of the American Fisheries Society, 122, 397403.Google Scholar
Varughese, M. M. (2011). A framework for modelling ecological communities and their interactions with the environment. Ecological Complexity, 8, 105112.Google Scholar
Vasseur, D. and Gaedke, U. (2007). Spectral analysis unmasks synchronous and compensatory dynamics in plankton communities. Ecology, 88, 20582071.Google Scholar
Winemiller, K. O. (1989). Patterns of variation in life history among South American fishes in seasonal environments. Oecologia, 81, 225241.Google Scholar
Winemiller, K. O. (1990). Spatial and temporal variation in tropical fish trophic networks. Ecological Monographs, 60, 331367.Google Scholar
Winemiller, K. O. (1992). Life history strategies and the effectiveness of sexual selection. Oikos, 62, 318327.Google Scholar
Winemiller, K. O. (2005). Life history strategies, population regulation, and their implications for fisheries management. Canadian Journal of Fisheries and Aquatic Sciences, 62, 872885.Google Scholar
Winemiller, K. O. and Rose, K. A. (1992). Patterns of life-history diversification in North American fishes: implications for population regulation. Canadian Journal of Fisheries and Aquatic Sciences, 49, 21962218.Google Scholar
Winemiller, K. O., Roelke, D. L., Cotner, J. B., et al. (2014). Top–down control of basal resources in a cyclically pulsing ecosystem. Ecological Monographs, 84, 621635.Google Scholar
Woodson, C. B., McManus, M. A., Tyburczy, J. A., et al. (2012). Coastal fronts set recruitment and connectivity patterns across multiple taxa. Limnology and Oceanography, 57, 582596.Google Scholar
Worm, B. and Duffy, J. E. (2003). Biodiversity, productivity and stability in real food webs. Trends in Ecology and Evolution, 18, 628632.Google Scholar
Yang, L. H. and Rudolf, V. H. W. (2010). Phenology, ontogeny and the effects of climate change on the timing of species interactions. Ecology Letters, 13, 110.Google Scholar
Yodzis, P. and Innes, S. (1992). Body size and consumer-resource dynamics. American Naturalist, 139, 11511175.Google Scholar
Zeglin, L. H., Bottomley, P. J., Jumpponen, A., et al. (2013). Altered precipitation regime affects the function and composition of soil microbial communities on multiple time scales. Ecology, 94, 23342345.Google Scholar
Zhou, C., Fujiwara, M., and Grant, W. E. (2013). Dynamics of a predator–prey interaction with seasonal reproduction and continuous predation. Ecological Modelling, 268, 2536.Google Scholar

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
×