Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-29T14:28:20.137Z Has data issue: false hasContentIssue false

Part I - Food Webs: Complexity and Stability

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. 7 - 104
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

Allesina, S. (2009). Cycling and cycling indices. In Ecosystem Ecology, ed. Jorgensen, S. E., Amsterdam, Elsevier, pp. 5057.Google Scholar
Allesina, S. and Tang, S. (2012). Stability criteria for complex ecosystems. Nature, 483, 205208.Google Scholar
Banašek-Richter, C., Cattin, M.-F., and Bersier, L.-F. (2004). Sampling effects and the robustness of quantitative and qualitative food-web descriptors. Journal of Theoretical Biology, 226, 2332.CrossRefGoogle ScholarPubMed
Bastolla, U., Fortuna, M. A., Pascual-García, A., et al. (2009). The architecture of mutualistic networks minimizes competition and increases biodiversity. Nature, 458, 10181021.Google Scholar
Benadi, G., Blüthgen, N., Hovestadt, T., and Poethke, H. J. (2012). Population dynamics of plant and pollinator communities: Stability reconsidered. The American Naturalist, 179, 157268.CrossRefGoogle ScholarPubMed
Blüthgen, N. (2010). Why network analysis is often disconnected from community ecology: A critique and an ecologist’s guide. Basic and Applied Ecology, 11, 185195.Google Scholar
Blüthgen, N., Fründ, J., Vázquez, D. P., and Menzel, F. (2008). What do interaction network metrics tell us about specialization and biological traits? Ecology, 89, 33873399.CrossRefGoogle ScholarPubMed
Boit, A., Martinez, N. D., Williams, R. J., and Gaedke, U. (2012). Mechanistic theory and modelling of complex food-web dynamics in Lake Constance. Ecology Letters, 15, 594602.Google Scholar
Case, T. J. (2000). An Illustrated Guide to Theoretical Ecology. Oxford, UK: Oxford University Press.Google Scholar
Cazelles, K., Araújo, M. B., Mouquet, N., and Gravel, D. (2016). A theory for species co-occurrence in interaction networks. Theoretical Ecology, 9, 3948.Google Scholar
Cohen, J. E. (1978). Food Webs and Niche Space. Princeton, NJ: Princeton University Press.Google ScholarPubMed
Darwin, C. R. (1859). The Origin of Species by Means of Natural Selection or the Preservation of Favoured Races in the Struggle for Life. London: J. Murray.Google Scholar
Dormann, C. F., Blüthgen, N., Fründ, J., and Gruber, B. (2009). Indices, graphs and null models: Analyzing bipartite ecological networks. Open Ecology Journal, 2, 724.Google Scholar
Fortuna, M. A., Krishna, A., and Bascompte, J. (2013). Habitat loss and the disassembly of mutalistic networks. Oikos, 122(6), 938942.CrossRefGoogle Scholar
Fründ, J., Dormann, C. F., and Tscharntke, T. (2011). Linné’s floral clock is slow without pollinators: Flower closure and plant–pollinator interaction webs. Ecology Letters, 14, 896904.Google Scholar
Fussmann, G. F. and Heber, G. (2002). Food web complexity and chaotic population dynamics. Ecology Letters, 5, 394401.Google Scholar
Goldwasser, L. and Roughgarden, J. (1997). Sampling effects and the estimation of food-web properties. Ecology, 78, 4154.CrossRefGoogle Scholar
Ings, T. C., Montoya, J. M., Bascompte, J., et al. (2009). Ecological networks: Beyond food webs. Journal of Animal Ecology, 78, 253269.Google Scholar
Ives, A. R. and Cardinale, B. J. (2004). Food-web interactions govern the resistance of communities after non-random extinctions. Nature, 429, 174177.Google Scholar
James, A., Pitchford, J. W., and Plank, M. J. (2012). Disentangling nestedness from models of ecological complexity. Nature, 487, 227230.Google Scholar
Joppa, L. N., Bascompte, J., Montoya, J. M., et al. (2009). Reciprocal specialization in ecological networks. Ecology Letters, 12, 961969.Google Scholar
Jordano, P. (1987). Patterns of mutualistic interactions in pollination and seed dispersal: Connectance, dependence asymmetries, and coevolution. American Naturalist, 129, 657677.Google Scholar
Kéfi, S., Berlow, E. L., Wieters, E., et al. (2012). More than a meal… Integrating non-feeding interactions into food webs. Ecology Letters, 15, 291300.Google Scholar
Martinez, N. D., Hawkins, B., Dawah, H. A., and Feifarek, B. P. (1999). Effects of sampling effort on characterization of food-web structure. Ecology, 80, 10441055.CrossRefGoogle Scholar
May, R. M. (1973). Stability and Complexity in Model Ecosystems. Princeton, NJ: Princeton University Press.Google Scholar
Melian, C. J. and Bascompte, J. (2002). Food web structure and habitat loss. Ecology Letters, 5, 3746.Google Scholar
Montoya, J. M., Pimm, S. L., and Solé, R. V. (2006). Ecological networks and their fragility. Nature, 442, 259264.Google Scholar
Morales-Castilla, I., Matias, M. G., Gravel, D., and Araújo, M. B. (2015). Inferring biotic interactions from proxies. Trends in Ecology and Evolution, 30, 347356.CrossRefGoogle ScholarPubMed
Morris, R. J., Gripenberg, S., Lewis, O. T., and Roslin, T. (2014). Antagonistic interaction networks are structured independently of latitude and host guild. Ecology Letters, 17, 340349.Google Scholar
Murdoch, W. W., Kendall, B. E., Nisbet, R. M., et al. (2002). Single-species models for many-species food webs. Nature, 417, 541543.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.CrossRefGoogle ScholarPubMed
Nielsen, A. and Bascompte, J. (2007). Ecological networks, nestedness and sampling effort. Journal of Ecology, 95, 11341141.Google Scholar
Odum, E. P. (1953). Fundamentals of Ecology. Philadelphia, PA: Saunders.Google Scholar
Olesen, J. M., Bascompte, J., Elberling, H., and Jordano, P. (2008). Temporal dynamics in a pollination network. Ecology, 89, 15731582.Google Scholar
Pimm, S. L. (1982). Food Webs. Chicago: Chicago University Press.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
Reuman, D. C., Mulder, C., Raffaelli, D., and Cohen, J. E. (2008). Three allometric relations of population density to body mass: Theoretical integration and empirical tests in 149 food webs. Ecology Letters, 11, 12161228.CrossRefGoogle ScholarPubMed
Rossberg, A. G. (2013). Food Webs and Biodiversity: Foundations, Models, Data. Oxford, UK: Wiley.Google Scholar
Schleuning, M., Fründ, J., Klein, A.-M., et al. (2012). Specialization of mutualistic interaction networks decreases toward tropical latitudes. Current Biology, 22, 17.Google Scholar
Solé, R. V., Alonso, D., and McKane, A. (2002). Self-organized instability in complex ecosystems. Philosophical Transactions of the Royal Society B: Biological Sciences, 357, 667681.Google Scholar
Sørensen, P. B., Damgaard, C. F., Strandberg, B., et al. (2011). A method for under-sampled ecological network data analysis: Plant-pollination as case study. Journal of Pollination Ecology, 6, 129139.Google Scholar
Suweis, S., Simini, F., Banavar, J. R., and Maritan, A. (2013). Emergence of structural and dynamical properties of ecological mutualistic networks. Nature, 500, 449452.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
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.Google Scholar
Vázquez, D. P. and Aizen, M. A. (2003). Null model analyses of specialization in plant–pollinator interactions. Ecology, 84, 24932501.Google Scholar
Vázquez, D. P. and Aizen, M. A. (2006). Community-wide patterns of specialization in plant–pollinator interactions revealed by null models. In Plant–Pollinator Interactions: From Specialization to Generalization, ed. Waser, N. M. and Ollerton, J.. Chicago, IL: University of Chicago Press, pp. 200219.Google Scholar
Vázquez, D. P., Morris, W. F., and Jordano, P. (2005). Interaction frequency as a surrogate for the total effect of animal mutualists on plants. Ecology Letters, 8, 10881094.Google Scholar
Vázquez, D. P., Chacoff, N., and Cagnolo, L. (2009). Evaluating multiple determinants of the structure of plant–animal mutualistic networks. Ecology, 90, 20392046.Google Scholar
Verdú, M. and Valiente-Banuet, A. (2011). The relative contribution of abundance and phylogeny to the structure of plant facilitation networks. Oikos, 120, 13511356.Google Scholar
Weiner, C. N., Werner, M., Linsenmair, K., and Blüthgen, N. (2014). Land use impacts on mutualistic networks: Disproportional declines in specialized pollinators via changes in flower composition. Ecology, 95, 466474.Google Scholar
Wells, K. and O’Hara, R. B. (2014). Species interactions: estimating per-individual interaction strength and covariates before simplifying data into per-species ecological networks. Methods in Ecology and Evolution, 4, 18.Google Scholar
Wells, K., Feldhaar, H., and O’Hara, R. B. (2014). Population fluctuations affect inference in ecological networks of multi-species interactions. Oikos, 123, 589598.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
Winfree, R., Williams, N. M., Dushoff, J., and Kremen, C. (2007). Native bees provide insurance against ongoing honey bee losses. Ecology Letters, 10, 11051113.Google Scholar

References

Allesina, S. and Tang, S. (2012). Stability criteria for complex ecosystems. Nature, 483, 205208.CrossRefGoogle ScholarPubMed
Bascompte, J. and Jordano, P. (2013). Mutualistic Networks. Princeton, NJ: Princeton University Press.Google Scholar
Burkholder, P. R. (1952). Cooperation and conflict among primitive organisms. American Scientist, 40, 601631.Google Scholar
Chen, X. and Cohen, J. E. (2001). Transient dynamics and food-web complexity in the Lotka–Volterra cascade model. Proceedings of the Royal Society B: Biological Sciences, 268, 869877.Google Scholar
de Ruiter, P. C., Wolters, V., and Moore, J. C. (2005). Dynamic Food Webs: Multispecies Assemblages, Ecosystem Development, and Environmental Change. San Diego, CA: Academic Press.Google Scholar
Emmerson, M. and Raffaelli, D. (2004). Body size, patterns of interaction strength and the stability of a real food web. Journal of Animal Ecology, 73, 399409.Google Scholar
Gause, G. F. (1932). Experimental studies on the struggle for existence. Journal of Experimental Biology, 9, 389402.CrossRefGoogle Scholar
Georgelin, E. and Loeuille, N. (2014). Dynamics of coupled mutualistic and antagonistic interactions, and their implications for ecosystem management. Journal of Theoretical Biology, 346, 6774.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. Ecological Monograph, 75, 335.CrossRefGoogle Scholar
Hutchinson, G. E. (1959). Homage to Santa Rosalia, or Why are there so many kinds of animals? American Naturalist, 93, 145159.Google Scholar
Johnson, N. C., Graham, J. H., and Smith, F. A. (1997). Functioning of mycorrhizal associations along the mutualism–parasitism continuum. New Phytologist, 135, 575585.Google Scholar
Kéfi, S., Berlow, E., Wieters, E., Navarrete, S., et al. (2012). More than a meal… Integrating non-feeding interactions into food webs. Ecology Letters, 15, 291300.Google Scholar
Kondoh, M. (2003). Foraging adaptation and the relationship between food-web complexity and stability. Science, 299, 13881391.Google Scholar
Kondoh, M. (2008). Building trophic modules into a persistent food web. Proceedings of the National Academy of Sciences of the United States of America, 105, 1663116635.Google Scholar
Lawlor, L. R. (1978). A comment on randomly constructed model ecosystems. American Naturalist, 112, 445447.Google Scholar
May, R. M. (1972). Will a large complex system be stable? Nature, 238, 413414.Google Scholar
Mélian, C. J., Bascompte, J., Jordano, P., and Křivan, V. (2009). Diversity in a complex ecological network with two interaction types. Oikos, 118, 122130.Google Scholar
Mougi, A. and Kondoh, M. (2012). Diversity of interaction types and ecological community stability. Science, 337, 349351.Google Scholar
Mougi, A. and Kondoh, M. (2014a). Adaptation in a hybrid world with multiple interaction types: A new mechanism for species coexistence. Ecological Research, 29, 113119.Google Scholar
Mougi, A. and Kondoh, M. (2014b). Stability of competition–antagonism–mutualism hybrid community and the role of community network structure. Journal of Theoretical Ecology, 360, 5458.Google ScholarPubMed
Mougi, A. and Kondoh, M. (2014c). Instability of a hybrid module of antagonistic and mutualistic interactions. Population Ecology, 56, 257263.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
Ohgushi, T. (2005). Indirect interaction webs: Herbivore-induced effects through trait change in plants. Annual Review of Ecology, Evolution, and Systematics, 36, 81105.Google Scholar
Ohgushi, T., Craig, T. P., and Price, P. W. (2007). Ecological Communities: Plant Mediation in Indirect Interaction Webs. Cambridge, UK: Cambridge University Press.Google Scholar
Ohgushi, T., Schmitz, O., and Holt, R. D. (2013). Trait-Mediated Indirect Interactions: Ecological and Evolutionary Perspectives. Cambridge, UK: Cambridge University Press.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
Ramos-Jiliberto, R., Valdovinos, F. S., Moisset de Espanés, P., and Flores, J. D. (2012). Topological plasticity increases robustness of mutualistic networks. Journal of Animal Ecology, 81, 896904.Google Scholar
Rúa, M. A. and Umbanhowar, J. (2015). Resource availability determines stability for mutualist–pathogen–host interactions. Theoretical Ecology, 8, 133148.Google Scholar
Sanders, D., Jones, C. G., Thébault, E., et al. (2014). Integrating ecosystem engineering and food webs. Oikos, 123, 513524.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
Suweis, S., Grilli, J., and Martian, A. (2014). Disentangling the effect of hybrid interactions and of the constant effort hypothesis on ecological community stability. Oikos, 123, 525532.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
Thompson, J. N. (2005). The Geographic Mosaic of Coevolution. Chicago, IL: University of Chicago Press.Google Scholar
Toju, H., Guimarães, P. R., Olesen, J. M., and Thompson, J. N. (2014). Assembly of complex plant–fungus networks. Nature Communications, 5, 5273.Google Scholar
Werner, E. E. and Peacor, S. D. (2003). A review of trait-mediated indirect interactions in ecological communities. Ecology, 84, 10831100.Google Scholar

References

Allesina, S. and Tang, S. (2012). Stability criteria for complex ecosystems. Nature, 483, 205208.Google Scholar
Banašek-Richter, C., Cattin, M. F., and Bersier, L.-F. (2004). Sampling effects and the robustness of quantitative and qualitative food-web descriptors. Journal of Theoretical Biology, 226, 2332.Google Scholar
Banašek-Richter, C., Bersier, L.-F., Cattin, M.-F., et al. (2009). Complexity in quantitative food webs. Ecology, 90, 14701477.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.CrossRefGoogle Scholar
Bersier, L. F., Banašek-Richter, C., and Cattin, M. F. (2002). Quantitative descriptors of food-web matrices. Ecology, 83, 23942407.Google Scholar
Brose, U., Williams, R. J., and Martinez, N. D. (2006). Allometric scaling enhances stability in complex food webs. Ecology Letters, 9, 12281236.Google Scholar
Dambacher, J. M., Luh, H. K., Li, H. W., and Rossignol, P. A. (2003). Qualitative stability and ambiguity in model ecosystems. American Naturalist, 161, 876888.Google Scholar
DeAngelis, D. L. (1975). Stability and connectance in food web models. Ecology, 56, 238243.Google Scholar
DeAngelis, D. L. (1992). Dynamics of Nutrient Cycling and Food Webs. New York: Chapman and Hall.Google Scholar
de Ruiter, P. C., Neutel, A. M., and Moore, J. C. (1995). Energetics, patterns of interaction strengths, and stability in real ecosystems. Science, 269, 12571260.Google Scholar
Drossel, B., McKane, A. J., and Quince, C. (2004). The impact of nonlinear functional responses on the long-term evolution of food web structure. Journal of Theoretical Biology, 229, 539548.CrossRefGoogle ScholarPubMed
Elton, C. (1927). Animal Ecology. New York: Macmillan.Google Scholar
Emmerson, M. C. and Raffaelli, D. (2004). Predator prey body size, interaction strength and the stability of a real food web. Journal of Animal Ecology, 73, 399409.Google Scholar
Gardner, M. R. and Ashby, W. R. (1970). Connectance of large dynamic (cybernetic) systems: critical values for stability. Nature, 228, 784.Google Scholar
Grime, J. P. (1977). Evidence for the existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. American Naturalist, 111, 11691194.Google Scholar
Grime, J. P. and Pierce, S. (2012). The Evolutionary Strategies That Shape Ecosystems. Oxford, UK: Wiley-Blackwell.Google Scholar
Grimm, V. and Wissel, C. (1997). Babel, or the ecological stability discussions: an inventory and analysis of terminology and a guide for avoiding confusion. Oecologia, 109, 323334.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
Hastings, H. M. (1982). The May–Wigner stability theorem. Journal of Theoretical Biology, 97, 155166.Google Scholar
Jacquet, C., Moritz, C., Morissette, L., et al. (2013). No complexity-stability relationship in natural communities. q-bio, arXiv:1307.5364.Google Scholar
James, A., Plank, M. J., Rossberg, A., et al. (2015). Constructing random matrices to represent real ecosystems. American Naturalist, 185, 680692.Google Scholar
Kondoh, M. (2003). Foraging adaptation and the relationship between food-web complexity and stability. Science, 299, 13881391.Google Scholar
Levins, R. (1968). Evolution in Changing Environments. Princeton, NJ: Princeton University Press.Google Scholar
Levins, R. (1974). The qualitative analysis of partially specified systems. Annals of the New York Academy of Science, 231, 123138.Google Scholar
Levins, R. (1977). Quantitative analysis of complex systems. In Mathematics and the Life Sciences, ed. Matthews, D. E., Vol. 18 of Lecture Notes in Biomathematics. New York: SpringerVerlag, pp. 152199.Google Scholar
Lindeman, R. L. (1942). The trophic-dynamic aspect of ecology. Ecology, 23, 399417.Google Scholar
MacArthur, R. (1955). Fluctuations of animal populations, and a measure of community stability. Ecology, 36, 533536.Google Scholar
Maruyama, M. (1963). The second cybernetics: deviation-amplifying mutual causal processes. American Scientist, 51, 164179.Google Scholar
May, R. M. (1972). Will a large complex system be stable? Nature, 238, 413414.Google Scholar
May, R. M. (1973). Stability and Complexity in Model Ecosystems. Princeton, NJ: Princeton University Press.Google Scholar
McCann, K. S. (2000). The diversity–stability debate. Nature, 405, 228233.Google Scholar
McCann, K., Hastings, A. G., and Huxel, R. (1998). Weak trophic interactions and the balance of nature. Nature, 395, 794798.Google Scholar
Menge, B. A. (1995). Indirect effects in marine rocky intertidal interaction webs: patterns and importance. Ecological Monographs, 65, 2174.Google Scholar
Mitchell, E. G. and Neutel, A. M. (2012). Feedback spectra of soil food webs across a complexity gradient, and the importance of three-species loops to stability. Theoretical Ecology, 5, 153159.Google Scholar
Montoya, J. M., Pimm, S. L., and Solé, R. V. (2006). Ecological networks and their fragility. Nature, 442, 259264.Google Scholar
Montoya, J. M., Woodward, G., Emmerson, M. C., and Solé, R. V. (2009). Press perturbations and indirect effects in real food webs. Ecology, 90, 24262433.Google Scholar
Moore, J. C., de Ruiter, P. C., and Hunt, H. W. (1993). Influence of productivity on the stability of real and model ecosystems. Science, 261, 906908.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
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
Neutel, A. M., Roerdink, J. B. T. M., and de Ruiter, P. C. (1994). Global stability of two-level detritus-decomposer food chains. Journal of Theoretical Biology, 171, 351353.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
Neutel, A. M., Heesterbeek, J. A. P., van de Koppel, J., et al. (2007). Reconciling complexity with stability in naturally assembling food webs. Nature, 449, 599602.Google Scholar
Novak, M., Wootton, J. T., Doak, D. F., et al. (2011). Predicting community responses to perturbations in the face of imperfect knowledge and network complexity. Ecology, 92, 836846.CrossRefGoogle ScholarPubMed
Odum, E. P. (1971). Fundamentals of Ecology, 3rd edn. Philadelphia, PA: Saunders.Google Scholar
Paine, R. T. (1988). Food webs: road maps of interactions or grist for theoretical development? Ecology, 69, 16481654.Google Scholar
Pimm, S. L. (1982). Food Webs. London, UK: Chapman & Hall.CrossRefGoogle Scholar
Pimm, S. L. and Lawton, J. H. (1977). Number of trophic levels in ecological communities. Nature, 268, 329331.Google Scholar
Pimm, S. L. and Lawton, J. H. (1978). On feeding on more than one trophic level. Nature, 275, 542544.Google Scholar
Polis, G. A. and Strong, D. R. (1996). Food web complexity and community dynamics. American Naturalist, 147, 813846.CrossRefGoogle Scholar
Rooney, N., McCann, K., Gellner, G., and Moore, J. C. (2006). Structural asymmetry and the stability of diverse food webs. Nature, 442, 265269.Google Scholar
Tang, S., Pawar, S., and Allesina, S. (2014). Correlation between interaction strengths drives stability in large ecological networks. Ecology Letters, 17, 10941100.Google Scholar
Ulanowicz, R. E., Holt, R. D., and Barfield, M. (2014). Limits on ecosystem trophic complexity: insights from ecological network analysis. Ecology Letters, 17, 127136.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
Wootton, J. T. (1994). Predicting direct and indirect effects: an integrated approach using experiments and path analysis. Ecology, 75, 151165.Google Scholar
Wootton, J. T. and Emmerson, M. (2005). Measurement of interaction strength in nature. Annual Review of Ecology Evolution and Systematics, 36, 419444.Google Scholar
Yodzis, P. (1981). The stability of real ecosystems. Nature, 289, 674676.Google Scholar

References

Abrams, P. (2009). When does greater mortality increase population size? The long history and diverse mechanisms underlying the hydra effect. Ecology Letters, 12, 462474.Google Scholar
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. and Tang, S. (2012). Stability criteria for complex ecosystems. Nature, 483, 205208.Google Scholar
Anderson, S., Kelly, D., Ladley, J., Molloy, S., and Terry, J. (2011). Cascading effects of bird functional extinction reduce pollination and plant density. Science, 331, 10681071.Google Scholar
Barnosky, A., Matzke, N., Tomiya, S., et al. (2011). Has the Earth’s sixth mass extinction already arrived? Nature, 471, 5157.Google Scholar
Baum, J. and Worm, B. (2009). Cascading top–down effects of changing oceanic predator abundances. Journal of Animal Ecology, 78, 699714.Google Scholar
Baum, J., Myers, R., Kehler, D., et al. (2003). Collapse and conservation of shark populations in the northwest Atlantic. Science, 299, 389392.Google Scholar
Beisner, B., Haydon, D., and Cuddington, K. (2003). Alternative stable states in ecology. Frontiers in Ecology and the Environment, 1, 376382.Google Scholar
Berg, S., Christianou, M., Jonsson, T., and Ebenman, B. (2011). Using sensitivity analysis to identify keystone species in size-based food webs. Oikos, 120, 510519.Google Scholar
Borrvall, C., Ebenman, B., and Jonsson, T. (2000). Biodiversity lessens the risk of cascading extinction in model food webs. Ecology Letters, 3, 131136.Google Scholar
Brose, U., Williams, R., and Martinez, N. (2006). Allometric scaling enhances stability in complex food webs. Ecology Letters, 9, 12281236.Google Scholar
Brummit, N., Bachman, S., and Moat, J. (2008). Applications of the IUCN Red List: towards a global barometer for plant diversity. Endangered Species Research, 6, 127135.Google Scholar
Butchart, S., Walpole, M., Collen, B., et al. (2010). Global biodiversity: indicators of recent declines. Science, 328, 11641168.Google Scholar
Casini, M., Hjelm, J., Molinero, J., 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
Cohen, J., Briand, F., and Newman, C. (1990). Community Food Webs: Data and Theory. Berlin Heidelberg: Springer-Verlag.Google Scholar
Collen, B., Loh, J., Whitmee, S., et al. (2009). Monitoring change in vertebrate abundance: the Living Planet Index. Conservation Biology, 23, 317327.Google Scholar
Colman, N., Gordon, C., Crowther, M., and Letnic, M. (2014). Lethal control of an apex predator has unintended cascading effects on forest mammal assemblages. Proceedings of the Royal Society B: Biological Sciences, 281, 20133094.Google Scholar
Colwell, R., Dunn, R., and Harris, N. (2012). Coextinction and persistence of dependent species in a changing world. Annual Review of Ecology, Evolution, and Systematics, 43, 183203.Google Scholar
Conner, R. (1988). Wildlife populations: minimally viable or ecologically functional? Wildlife Society Bulletin, 16, 8084.Google Scholar
Curtsdotter, A., Binzer, A., Brose, U., et al. (2011). Robustness to secondary extinctions: comparing trait-based sequential deletions in static and dynamic food webs. Basic and Applied Ecology, 12, 571580.Google Scholar
Cury, P., Boyd, I., Bonhommeau, S., et al. (2011). Global seabird response to forage fish depletion: one-third for the birds. Science, 334, 17031706.Google Scholar
de Roos, A. and Persson, L. (2013). Population and Community Ecology of Ontogenetic Development. Princeton, NJ: Princeton University Press.Google Scholar
Di Marco, M., Boitani, L., Mallon, D., et al. (2014). A retrospective evaluation of the global decline of carnivores and ungulates. Conservation Biology, DOI: 10.1111/cobi.12249.Google Scholar
Dirzo, R., Young, H., Galetti, M., et al. (2014). Defaunation in the Anthropocene. Science, 345, 401406.Google Scholar
Dunne, J., Williams, R., and Martinez, N. (2002). Network structure and biodiversity loss in food webs: robustness increases with connectance. Ecology Letters, 5, 558567.Google Scholar
Ebenman, B. (1987). Niche differences between age classes and intraspecific competition in age-structured populations. Journal of Theoretical Biology, 124, 2533.Google Scholar
Ebenman, B. (1992). Evolution in organisms that change their niches during the life cycle. American Naturalist, 139, 9901021.Google Scholar
Ebenman, B. and Jonsson, T. (2005). Using community viability analysis to identify fragile systems and keystone species. Trends in Ecology and Evolution, 20, 568575.Google Scholar
Ebenman, B. and Persson, L. (1988). Size-Structured Populations: Ecology and Evolution. Berlin Heidelberg: Springer Verlag.Google Scholar
Ebenman, B., Law, R., and Borrvall, C. (2004). Community viability analysis: the response of ecological communities to species loss. Ecology, 85, 25912600.Google Scholar
Eklöf, A. and Ebenman, B. (2006). Species loss and secondary extinctions in simple and complex model communities. Journal of Animal Ecology, 75, 239246.Google Scholar
Estes, J. A. and Palmisano, J. F. (1974) Sea otters: their role in structuring nearshore communities. Science, 185, 10581060.Google Scholar
Estes, J., Duggins, D., and Rathbun, G. (1989). The ecology of extinctions in kelp forest communities. Conservation Biology, 3, 252264.Google Scholar
Estes, J., Tinker, T., and Bodkin, J. (2010). Using ecological function to develop recovery criteria for depleted species: sea otters and kelp forests in the Aleutian archipelago. Conservation Biology, 24, 852860.Google Scholar
Estes, J. A., Terborgh, J., Brashares, J. S., et al. (2011). Trophic downgrading of planet Earth. Science, 333, 301306.Google Scholar
Fowler, M. (2010). Extinction cascades and the distribution of species interactions. Oikos, 119, 864873.Google Scholar
Frank, K., Petrie, B., Choi, J., and Leggett, W. (2005). Trophic cascades in a formerly cod-dominated ecosystem. Science, 308, 16211623.Google Scholar
Froyd, C., Coffey, E., van der Knapp, W., et al. (2014). The ecological consequences of megafaunal loss: giant tortoises and wetland biodiversity. Ecology Letters, 17, 144154.Google Scholar
Galetti, M., Guevara, R., Côrtes, M., et al. (2013). Functional extinction of birds drives rapid evolutionary changes in seed size. Science, 340, 10861090.Google Scholar
Gaston, K., Blackburn, T., and Goldewijk, K. (2003). Habitat conversion and global avian biodiversity loss. Proceedings of the Royal Society B: Biological Sciences, 270, 12931300.Google Scholar
Gilbert, B., Tunney, T., McCann, K., et al. (2014). A bioenergetics framework for the temperature dependence of trophic interactions. Ecology Letters, 17, 902914.Google Scholar
Gross, T., Rudolf, L., Levin, S., and Dieckmann, U. (2009). Generalized models reveal stabilizing factors in food webs. Science, 325, 747750.Google Scholar
IUCN (2014). IUCN Red List of Threatened Species. Version 2014.1.Google Scholar
Ives, A., Dennis, B., Cottingham, K., and Carpenter, S. (2003). Estimating community stability and ecological interactions from time-series data. Ecological Monographs, 73, 301330.Google Scholar
Jackson, J., Kirby, M., Berger, H., et al. (2001). Historical overfishing and the recent collapse of coastal ecosystems. Science, 293, 629638.Google Scholar
Joppa, L., Roberts, D., and Pimm, S. (2011). How many species of flowering plants are there? Proceedings of the Royal Society B: Biological Sciences, 278, 554559.Google Scholar
Kaneryd, L., Borrvall, C., Berg, S., et al. (2012). Species-rich ecosystems are vulnerable to cascading extinctions in an increasingly variable world. Ecology and Evolution, 2, 858874.Google Scholar
Lande, R. (1993). Risk of population extinction from demographic and environmental stochasticity and random catastrophes. American Naturalist, 142, 911927.Google Scholar
Lewis, H. and Law, R. (2007). Effects of dynamics on ecological networks. Journal of Theoretical Biology, 247, 6476.Google Scholar
Loh, J., Green, R., Rickets, T., et al. (2005). The Living Planet Index: using species population time series to track trends in biodiversity. Philosophical Transaction of the Royal Society B: Biological Sciences, 360, 289295.Google Scholar
Lundberg, P., Ranta, E., and Kaitala, V. (2000). Species loss leads to community closure. Ecology Letters, 3, 465468.Google Scholar
May, R. (1972). Will a large complex system be stable? Nature, 238, 413414.Google Scholar
May, R. (1977). Thresholds and breakpoints in ecosystems with a multiplicity of stable states. Nature, 269, 471477.Google Scholar
McCann, K. (2000). The diversity–stability debate. Nature, 405, 228233.Google Scholar
McCann, K., Rasmussen, J., Umbanhowar, J., and Chase, J. (2005). The dynamics of spatially coupled food webs. Ecology Letters, 8, 513523.Google Scholar
McConkey, K. and Drake, D. (2006). Flying foxes cease to function as seed dispersers long before they become rare. Ecology, 87, 271276.Google Scholar
McConkey, K. and O’Farrill, G. (2015). Cryptic function loss in animal populations. Trends in Ecology and Evolution, 30, 182189.Google Scholar
Miller, T. and Rudolf, V. (2011). Thinking inside the box: community-level consequences of stage-structured populations. Trends in Ecology and Evolution, 26, 457466.Google Scholar
Mougi, A. and Kondoh, M. (2013). Diversity of interaction types and ecological community stability. Science, 337, 349351.Google Scholar
Myers, R. and Worm, B. (2003). Rapid worldwide depletion of predatory fish communities. Nature, 423, 280283.Google Scholar
Neutel, A., Heesterbeek, J., and de Ruiter, P. (2002). Stability in real food webs: weak links in long loops. Science, 296, 11201123.Google Scholar
Novaro, A., Funes, M., and Walker, S. (2000). Ecological extinction of native prey of a carnivore assemblage in Argentine Patagonia. Biological Conservation, 92, 2533.Google Scholar
Pereira, H., Leadley, P., Proenca, V., et al. (2010). Scenarios for global biodiversity in the 21st century. Science, 330, 14961501.Google Scholar
Petchey, O., Eklöf, A., Borrvall, C., and Ebenman, B. (2008). Trophically unique species are vulnerable to cascading extinction. American Naturalist, 171, 568579.Google Scholar
Pimm, S. (1980). Food web design and the effects of species deletion. Oikos, 35, 139149.Google Scholar
Pimm, S., Jenkins, C., Abell, R., et al. (2014). The biodiversity of species and their rates of extinction, distribution, and protection. Science, 344, 1246752.Google Scholar
Redford, K. and Feinsinger, P. (2001). The half-empty forest: sustainable use and the ecology of interactions. In Conservation of Exploited Species, ed. Reynolds, D., Mace, G., Redfor, K., and Robinson, J.. Cambridge, UK: Cambridge University Press. pp. 370399.Google Scholar
Ripple, W., Estes, J., Bechta, R., et al. (2014). Status and ecological effects of the world’s largest carnivores. Science, 342, 1241484.Google Scholar
Rudolf, V. and Lafferty, K. (2011). Stage structure alters how complexity affects stability of ecological networks. Ecology Letters, 14, 7579.Google Scholar
Sabo, J. (2008). Population viability and species interactions: life outside the single-species vacuum. Biological Conservation, 14, 276286.Google Scholar
Sanders, D., Sutter, L., and van Veen, F. (2013). The loss of indirect interactions leads to cascading extinctions of carnivores. Ecology Letters, 16, 664669.Google Scholar
Sanders, D., Kehoe, R., and van Veen, F. (2015). Experimental evidence for the population-dynamic mechanisms underlying extinction cascades of carnivores. Current Biology, 25, 14.Google Scholar
Säterberg, T., Sellman, S., and Ebenman, B. (2013). High frequency of functional extinctions in ecological networks. Nature, 499, 468470.Google Scholar
Scheffer, M., Carpenter, S., Foley, J., Folke, C., and Walker, B. (2001). Catastrophic shifts in ecosystems. Nature, 413, 591596.Google Scholar
Schipper, J., Chanson, J., Chiozza, T., et al. (2008). The status of the world’s land and marine mammals: diversity, threat, and knowledge. Science, 322, 225230.Google Scholar
Sekercioglu, C., Daily, G., and Ehrlich, P. (2004). Ecosystem consequences of bird declines. Proceedings of the National Academy of Sciences of the United States of America, 101, 1804218047.Google Scholar
Sellman, S., Säterberg, T., and Ebenman, B. (2015) Pattern of functional extinctions in ecological networks with a variety of interaction types. Theoretical Ecology, DOI 10.1007/s12080-015–0275-7.Google Scholar
Shaffer, M. L. (1981). Minimum population sizes for species conservation. BioScience, 31, 131134.Google Scholar
Smith, A., Brown, C., Bulman, C., et al. (2011). Impact of fishing low-trophic level species on marine ecosystems. Science, 333, 11471150.Google Scholar
Solé, R. and Montoya, J. (2001). Complexity and fragility in ecological networks Proceedings of the Royal Society B: Biological Sciences, 268, 20392045.Google Scholar
Soulé, M., Estes, J., Berger, J., and Martinez Del Rio, C. (2003). Ecological effectiveness: conservation goals for interactive species. Conservation Biology, 17, 12381250.Google Scholar
Soulé, M., Estes, J., Miller, B., and Honnold, D. (2005). Strongly interacting species: conservation policy, management, and ethics. BioScience, 55, 168176.Google Scholar
Stouffer, D. 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
Thebault, E. and Fontaine, C. (2010). Stability of ecological communities and the architecture of mutualistic and trophic networks. Science, 329, 853856.Google Scholar
Thebault, E., Huber, V., and Loreau, M. (2007). Cascading extinctions and ecosystem functioning: contrasting effects of diversity depending on food web structure. Oikos, 116, 163173.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
Wilmers, C., Post, E., Peterson, R., and Vucetich, J. (2006). Predator disease out-break modulates top–down, bottom–up and climatic effects on herbivore population dynamics. Ecology Letters, 9, 383389.Google Scholar
Wollrab, S., Diehl, S., and de Roos, A. (2012). Simple rules describe bottom–up and top–down control in food webs with alternative energy pathways. Ecology Letters, 15, 935946.Google Scholar

References

Albrecht, M., Duelli, P., Schmid, B., and Muller, C. B. (2007). Interaction diversity within quantified insect food webs in restored and adjacent intensively managed meadows. Journal of Animal Ecology, 76, 10151025.Google Scholar
Altermatt, F. and Pearse, I. S. (2011). Similarity and specialization of the larval versus adult diet of European butterflies and moths. American Naturalist, 178, 372382.Google Scholar
Amundsen, P.-A., Lafferty, K. D., Knudsen, R. et al. (2009). Food web topology and parasites in the pelagic zone of a subarctic lake. Journal of Animal Ecology, 78, 563572.Google Scholar
Bascompte, J., Jordano, P., Melián, C. J., and Olesen, J. M. (2003). The nested assembly of plant–animal mutualistic networks. Proceedings of the National Academy of Sciences of the United States of America, 100, 93839387.Google Scholar
Bersier, L., Banašek-Richter, C., and Cattin, M. (2002). Quantitative descriptors of food-web matrices. Ecology, 83, 23942407.Google Scholar
Cagnolo, L., Salvo, A., and Valladares, G. (2011). Network topology: patterns and mechanisms in plant–herbivore and host–parasitoid food webs. Journal of Animal Ecology, 80, 342351.Google Scholar
Chen, H.-W., Liu, W.-C., Davis, A. J., et al. (2008). Network position of hosts in food webs and their parasite diversity. Oikos, 117, 18471855.Google Scholar
Elias, M., Fontaine, C., and van Veen, F. J. F. (2013). Evolutionary history and ecological processes shape a local multilevel antagonistic network. Current Biology, 23, 13551359.Google Scholar
Evans, D. M., Pocock, M. J. O., and Memmott, J. (2013). The robustness of a network of ecological networks to habitat loss. Ecology Letters, 16, 844852.Google Scholar
Fontaine, C., Guimarães, P. R., Kéfi, S., et al. (2011). The ecological and evolutionary implications of merging different types of networks. Ecology Letters, 14, 11701181.Google Scholar
Jordano, P., Bascompte, J., and Olesen, J. M. (2003). Invariant properties in coevolutionary networks of plant-animal interactions. Ecology Letters, 6, 6981.Google Scholar
Kéfi, S., Berlow, E. L., Wieters, E., et al. (2012). More than a meal… integrating non-feeding interactions into food webs. Ecology Letters, 15, 291300.Google Scholar
Krasnov, B. R., Fortuna, M. A., Mouillot, D., et al. (2012). Phylogenetic signal in module composition and species connectivity in compartmentalized host–parasite networks. American Naturalist, 179, 501511.Google Scholar
Leibold, M. (1996). A graphical model of keystone predators in food webs: trophic regulation of abundance, incidence, and diversity patterns in communities. American Naturalist, 147, 784812.Google Scholar
McCann, K., Hastings, A., and Huxel, G. (1998). Weak trophic interactions and the balance of nature. Nature, 395, 794798.Google Scholar
Melián, C. J., Bascompte, J., Jordano, P., and Krivan, V. (2009). Diversity in a complex ecological network with two interaction types. Oikos, 118, 122130.Google Scholar
Memmott, J., Waser, N. M., and Price, M. V. (2004). Tolerance of pollination networks to species extinctions. Proceedings of the Royal Society B: Biological Sciences, 271, 26052611.Google Scholar
Montesinos-Navarro, A., Segarra-Moragues, J. G., Valiente-Banuet, A., and Verdú, M. (2012). The network structure of plant–arbuscular mycorrhizal fungi. New Phytologist, 194, 536547.Google Scholar
Mougi, A. and Kondoh, M. (2012). Diversity of interaction types and ecological community stability. Science, 337, 349351.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
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
Sanders, D., Jones, C. G., Thébault, E., et al. (2014). Integrating ecosystem engineering and food webs. Oikos, 123, 513524.Google Scholar
Sauve, A. M. C., Fontaine, C., and Thébault, E. (2014). Structure–stability relationships in networks combining mutualistic and antagonistic interactions. Oikos, 123, 378384.Google Scholar
Sauve, A. M. C., Thébault, E., Pocock, M. J. O., and Fontaine, C. (2016a). How plants connect pollination and herbivory networks and their contribution to community stability. Ecology, 97, 908918.Google Scholar
Sauve, A. M. C., Fontaine, C., and Thébault, E. (2016b). Stability of a diamond-shaped module with multiple interaction types. Theoretical Ecology, 9, 2737.Google Scholar
Thébault, E. (2013). Identifying compartments in presence-absence matrices and bipartite networks: insights into modularity measures. Journal of Biogeography, 40, 759768.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
Van Veen, F. J. F., Müller, C. B., Pell, J. K., and Godfray, H. C. J. (2008). Food web structure of three guilds of natural enemies: predators, parasitoids and pathogens of aphids. Journal of Animal Ecology, 77, 191200.Google Scholar
Vazquez, D. P., Poulin, R., Krasnov, B. R., and Shenbrot, G. I. (2005). Species abundance and the distribution of specialization in host–parasite interaction networks. Journal of Animal Ecology, 74, 946955.Google Scholar
Verdú, M. and Valiente-Banuet, A. (2008). The nested assembly of plant facilitation networks prevents species extinctions. American Naturalist, 172, 751760.Google Scholar

References

Aguiar, M. R. and Sala, O. E. (1999). Patch structure, dynamics and implications for the functioning of arid ecosystems. Trends in Ecology and Evolution, 14(7), 273277.Google Scholar
Allesina, S. and Tang, S. (2012). Stability criteria for complex ecosystems. Nature, 483(7388), 205208.Google Scholar
Amarasekare, P. (2008). Spatial dynamics of food webs. Annual Review of Ecology, Evolution, and Systematics, 39(1), 479500.Google Scholar
Arditi, R., Michalski, J., and Hirzel, A. H. (2005). Rheagogies: modelling non-trophic effects in food webs. Ecological Complexity, 2(3), 249258.Google Scholar
Bascompte, J., Jordano, P., Melián, C. J., and Olesen, J. M. (2003). The nested assembly of plant–animal mutualistic networks. Proceedings of the National Academy of Sciences of the United States of America, 100(16), 93839387.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(3), 585598.Google Scholar
Blüthgen, N., Menzel, F., Hovestadt, T., Fiala, B., and Blüthgen, N. (2007). Specialization, constraints, and conflicting interests in mutualistic networks. Current Biology, 17(4), 341346.Google Scholar
Boccaletti, S., Bianconi, G., Criado, R., et al. (2014). The structure and dynamics of multilayer networks. Physics Reports, 544(1), 1122.Google Scholar
Borer, E. T., Anderson, K., Blanchette, C. A., et al. (2002). Topological approaches to food web analyses: a few modifications may improve our insights. Oikos, 99(2), 397401.Google Scholar
Brose, U., Cushing, L., Berlow, E. L., et al. (2005). Body sizes of consumers and their resources. Ecology, 86(9), 2545.Google Scholar
Brose, U., Jonsson, T., Berlow, E. L., et al. (2006). Consumer–resource body-size relationships in natural food webs. Ecology, 87(10), 24112417.Google Scholar
Cardillo, A., Gómez-Gardeñes, J., Zanin, M., et al. (2013). Emergence of network features from multiplexity. Scientific Reports, 3, 1344.Google Scholar
Cohen, J. E., Pimm, S. L., Yodzis, P., and Saldana, J. (1993). Body sizes of animal predators and animal prey in food webs. Journal of Animal Ecology, 62(1), 6778.Google Scholar
Darwin, C. (1859). The Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life. London: John Murray, Albemarle Street.Google Scholar
Davis, A. J., Liu, W., Perner, J., and Voigt, W. (2004). Reliability characteristics of natural functional group interaction webs. Evolutionary Ecology Research, 6(8), 11451166.Google Scholar
de Ruiter, P. C., Neutel, A.-M., and Moore, J. C. (1995). Energetics, patterns of interaction strengths, and stability in real ecosystems. Science, 269(5228), 12571260.Google Scholar
Donohue, I., Petchey, O. L., Montoya, J. M., et al. (2013). On the dimensionality of ecological stability. Ecology Letters, 16, 421429.Google Scholar
Fontaine, C., Guimarães, P. R., Kéfi, S., et al. (2011). The ecological and evolutionary implications of merging different types of networks. Ecology Letters, 14(11), 11701181.Google Scholar
Genini, J., Morellato, L. P. C., Guimarães, P. R., and Olesen, J. M. (2010). Cheaters in mutualism networks. Biology Letters, rsbl20091021.Google Scholar
Goudard, A. and Loreau, M. (2008). Nontrophic interactions, biodiversity, and ecosystem functioning: an interaction web model. American Naturalist, 171(1), 91106.Google Scholar
Gross, K. (2008). Positive interactions among competitors can produce species-rich communities. Ecology Letters, 11(9), 929936.Google Scholar
Holt, R. D. (2002). Food webs in space: on the interplay of dynamic instability and spatial processes. Ecological Research, 17(2), 261273.Google Scholar
Ings, T. C., Montoya, J. M., Bascompte, J., et al. (2009). Ecological networks: beyond food webs. Journal of Animal Ecology, 78(1), 253269.Google Scholar
Ives, A. R. and Carpenter, S. R. (2007). Stability and diversity of ecosystems. Science, 317(5834), 5862.Google Scholar
Jordán, F., Lauria, M., Scotti, M., et al. (2015). Diversity of key players in the microbial ecosystems of the human body. Scientific Reports, 5, 15920.Google Scholar
Jordano, P., Bascompte, J., and Olesen, J. M. (2003). Invariant properties in coevolutionary networks of plant–animal interactions. Ecology Letters, 6(1), 6981.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
Kéfi, S., Berlow, E. L., Wieters, E. A., et al. (2015). Network structure beyond food webs: mapping non-trophic and trophic interactions on Chilean rocky shores. Ecology, 96(1), 291303.Google Scholar
Kéfi, S., Miele, V., Wieters, E. A., Navarrete, S. A., and Berlow, E. L. (2016). How structured is the entangled bank? The surprisingly simple organization of multiplex ecological networks leads to increased persistence and resilience. PLoS Biology, 14(8), e1002527.Google Scholar
Kivelä, M., Arenas, A., Barthelemy, M., et al. (2014). Multilayer networks. Journal of Complex Networks, 2(3), 203271.Google Scholar
Krasnov, B. R., Fortuna, M. A., Mouillot, D., et al. (2012). Phylogenetic signal in module composition and species connectivity in compartmentalized host–parasite networks. American Naturalist, 179(4), 501511.Google Scholar
Lafferty, K. D., Dobson, A. P., and Kuris, A. M. (2006). Parasites dominate food web links. Proceedings of the National Academy of Sciences of the United States of America, 103(30), 1121111216.Google Scholar
Lafferty, K. D., Allesina, S., Arim, M., et al. (2008). Parasites in food webs: the ultimate missing links. Ecology Letters, 11(6), 533546.Google Scholar
Lin, Y. and Sutherland, W. J. (2013). Color and degree of interspecific synchrony of environmental noise affect the variability of complex ecological networks. Ecological Modelling, 263, 162173.Google Scholar
Lurgi, M., Montoya, D., and Montoya, J. M. (2016). The effects of space and diversity of interaction types on the stability of complex ecological networks. Theoretical Ecology, 9(1), 313.Google Scholar
May, R. M. (1972). Will a large complex system be stable? Nature, 238(5364), 413414.Google Scholar
McCann, K. S., Rasmussen, J. B., and Umbanhowar, J. (2005). The dynamics of spatially coupled food webs. Ecology Letters, 8(5), 513523.Google Scholar
Melián, C. J., Bascompte, J., Jordano, P., and Krivan, V. (2009). Diversity in a complex ecological network with two interaction types. Oikos, 118(1), 122130.Google Scholar
Montoya, J. M., Pimm, S. L., and Solé, R. V. (2006). Ecological networks and their fragility. Nature, 442(7100), 259264.Google Scholar
Mougi, A. and Kondoh, M. (2012). Diversity of interaction types and ecological community stability. Science, 337(6092), 349351.Google Scholar
Mucha, P. J., Richardson, T., Macon, K., Porter, M. A., and Onnela, J.-P. (2010). Community structure in time-dependent, multiscale, and multiplex networks. Science, 328(5980), 876878.Google Scholar
Neutel, A.-M., Heesterbeek, J. A. P., van de Koppel, J., et al. (2007). Reconciling complexity with stability in naturally assembling food webs. Nature, 449(7162), 599602.Google Scholar
Ohgushi, T., Schmitz, O., and Holt, R. D. (2012). Trait-Mediated Indirect Interactions. Cambridge, UK: Cambridge University Press.Google Scholar
Okuyama, T. and Holland, J. N. (2008). Network structural properties mediate the stability of mutualistic communities. Ecology Letters, 11(3), 208216.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(1524), 17551779.Google Scholar
Pilosof, S., Porter, M. A., Pascual, M., and Kéfi, S. (2017). The multilayer nature of ecological networks. Nature in Ecology and Evolution, 1, 101.Google Scholar
Pimm, S. L. (1982). Food Webs. Chicago, IL: University of Chicago Press.Google Scholar
Pimm, S. L. (1984). The complexity and stability of ecosystems. Nature, 307(5949), 321326.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(6071), 973977.Google Scholar
Saiz, H., and Alados, C. L. (2011). Effect of Stipa tenacissima L. on the structure of plant co-occurrence networks in a semi-arid community. Ecological Research, 26(3), 595603.Google Scholar
Sander, E. L., Wootton, J. T., and Allesina, S. (2015). What can interaction webs tell us about species roles? PLOS Computational Biology, 11(7), e1004330.Google Scholar
Sanders, D. and van Veen, F. J. F. (2011). Ecosystem engineering and predation: the multi-trophic impact of two ant species. Journal of Animal Ecology, 80, 569576.Google Scholar
Sanders, D., Jones, C. G., Thébault, E., et al. (2014). Integrating ecosystem engineering and food webs. Oikos, 123(5), 513524.Google Scholar
Sauve, A. M. C., Fontaine, C., and Thébault, E. (2014). Structure–stability relationships in networks combining mutualistic and antagonistic interactions. Oikos, 123(3), 378384.Google Scholar
Sauve, A. M. C., Thébault, E., Pocock, M. J. O., and Fontaine, C. (2016). How plants connect pollination and herbivory networks and their contribution to community stability. Ecology, 97(4), 908917.Google Scholar
Soliveres, S., Smit, C., and Maestre, F. T. (2015). Moving forward on facilitation research: response to changing environments and effects on the diversity, functioning, and evolution of plant communities. Biological Reviews of the Cambridge Philosophical Society, 90(1), 297313.Google Scholar
Thébault, E. and Fontaine, C. (2010). Stability of ecological communities and the architecture of mutualistic and trophic networks. Science, 329(5993), 853856.Google Scholar
Vázquez, D. P., Poulin, R., Krasnov, B. R., and Shenbrot, G. I. (2005). Species abundance and the distribution of specialization in host–parasite interaction networks. Journal of Animal Ecology, 74(5), 946955.Google Scholar
Verdú, M., and Valiente-Banuet, A. (2008). The nested assembly of plant facilitation networks prevents species extinctions. American Naturalist, 172(6), 751760.Google Scholar
Williams, R. J. and Martinez, N. D. (2000). Simple rules yield complex food webs. Nature, 404(6774), 180183.Google Scholar
Wood, S. A., Lilley, S. A., Schiel, D. R., and Shurin, J. B. (2010). Organismal traits are more important than environment for species interactions in the intertidal zone. Ecology Letters, 13(9), 11601171.Google Scholar
Wootton, J. T., Sander, E. L., and Allesina, S. (2015). Data from: What can interaction webs tell us about species roles? Dryad Digital Repository, http://dx.doi.org/10.5061/dryad.39jv1 [Accessed March 22, 2017].Google Scholar
Yodzis, P. (1981). The stability of real ecosystems. Nature, 289, 674676.Google Scholar

References

Amarasekare, P. (2008). Spatial dynamics of food webs. Annual Review of Ecology, Evolution, and Systematics, 39(1), 479500.Google Scholar
Archibald, S., Staver, A. C., and Levin, S. A. (2012). Evolution of human-driven fire regimes in Africa. Proceedings of the National Academy of Sciences of the United States of America, 109, 847852.Google Scholar
Bellard, C., Bertelsmeier, C., Leadley, P., Thuiller, W., and Courchamp, F. (2012). Impacts of climate change on the future of biodiversity. Ecology Letters, 15, 365377.Google Scholar
Berlow, E. L. (1999). Strong effects of weak interactions in ecological communities. Nature, 398, 330334.Google Scholar
Bolnick, D. I., Amarasekare, P., Araújo, M. S., et al. (2011). Why intraspecific trait variation matters in community ecology. Trends in Ecology and Evolution, 26(4), 183192.Google Scholar
Chesson, P. and Huntly, N. (1997). The roles of harsh and fluctuating conditions in the dynamics of ecological communities. American Naturalist, 150, 519553.Google Scholar
Clapcott, J. E. and Barmuta, L. A. (2010). Forest clearance increases metabolism and organic matter processes in small headwater streams. Journal of the North American Benthological Society, 29, 546561.Google Scholar
Closs, G. P. and Lake, P. S. (1994). Spatial and temporal variation in the structure of an intermittent-stream food web. Ecological Monographs, 64, 121.Google Scholar
DeAngelis, D. L. (1975). Stability and connectance in food web models. Ecology, 56, 238243.Google Scholar
De Lange, H. J., Kramer, K., and Faber, J. H. (2013). Two approaches using traits to assess ecological resilience: a case study on earthworm communities. Basic and Applied Ecology, 14, 6473.Google Scholar
Doak, D. F., Bigger, D., Harding, E. K., et al. (1998). The statistical inevitability of stability–diversity relationships in community ecology. American Naturalist, 151, 264276.Google Scholar
Dunne, J. A. (2006). The network structure of food webs. In Ecological Networks: Linking Structure to Dynamics in Food Webs, ed. Pascual, M. and Dunne, J. A., Oxford, UK: Oxford University Press, pp. 2786.Google Scholar
Dunne, J. A., Williams, R. J., and Martinez, N. D. (2002). Network structure and biodiversity loss in food webs: robustness increases with connectance. Ecology Letters, 5, 558567.Google Scholar
Ehrlich, P. R. and Ehrlich, A. H. (1983). Extinction: The Causes and Consequences of the Disappearance of Species. New York: Random House.Google Scholar
Fagan, W. F. (1997). Omnivory as a stabilizing feature of natural communities. American Naturalist, 150, 554567.Google Scholar
Gilbert, A. J. (2009). Connectance indicates the robustness of food webs when subjected to species loss. Ecological Indicators, 9, 7280.Google Scholar
Göthe, E., Lepori, F., and Malmqvist, B. (2009). Forestry affects food webs in northern Swedish coastal streams. Fundamental and Applied Limnology/Archiv Für Hydrobiologie, 175, 281294.Google Scholar
Growns, I. and Davis, J. (1994). Effects of forestry activities (clearfelling) on stream macroinvertebrate fauna in south-western Australia. Marine and Freshwater Research, 45, 963975.Google Scholar
Harding, J. S., Benfield, E. F., Bolstad, P. V., Helfman, G. S., and Jones, E. B. (1998). Stream biodiversity: the ghost of land use past. Proceedings of the National Academy of Sciences of the United States of America, 95, 1484314847.Google Scholar
Holling, C. S. (1973). Resilience and stability of ecological systems. Annual Review of Ecology and Systematics, 4, 123.Google Scholar
Hooper, D. U., Chapin, F. S. I., Ewel, J. J., et al. (2005). Effects of biodiversity on ecosystem functioning: a consensus of current knowledge. Ecological Monographs, 75, 335.Google Scholar
Lancaster, A. L. and Robertson, J. R. (1995). Microcrustacean prey and macroinvertebrate predators in a stream food web. Freshwater Biology, 34, 123134.Google Scholar
Levin, S. A. (1970). Community equilibria and stability, and an extension of the competitive exclusion principle. American Naturalist, 104, 413423.Google Scholar
Lindeman, R. L. (1942). The trophic dynamics aspect of ecology. Ecology, 23, 399418.Google Scholar
Lubetkin, S. C. and Simenstad, C. A. (2004). Multi-source mixing models to quantify food web sources and pathways. Journal of Applied Ecology, 41, 9961008.Google Scholar
Martinez, N. D., Williams, R. J., Berlow, E. L., and Dunne, J. A. (2002). Two degrees of separation in complex food webs. Proceedings of the National Academy of Sciences of the United States of America, 99, 1291312916.Google Scholar
May, R. M. (1971). Stability in multispecies community models. Mathematical Biosciences, 12, 5979.Google Scholar
May, R. M. (1972). Will a large complex system be stable? Nature, 238, 413414.Google Scholar
McCann, K. S. (2000). The diversity–stability debate, Nature, 405, 227233.Google Scholar
McCann, K., Hastings, A., and Huxel, G. (1998). Weak trophic interactions and the balance of nature. Nature, 395, 794798.Google Scholar
McCann, K. S., Rasmussen, J. B., and Umbanhowar, J. (2005). The dynamics of spatially coupled food webs. Ecology Letters, 8, 513523.Google Scholar
Memmott, J., Martinez, N. D., and Cohen, J. E. (2000). Predators, parasitoids and pathogens: species richness, trophic generality and body sizes in a natural food web. Journal of Animal Ecology, 69, 115.Google Scholar
Menge, B. A. (1995). Indirect effects in marine rocky intertidal interaction webs: patterns and importance. Ecological Monographs, 65, 2174.Google Scholar
Morán-Ordóñez, A., Pavlova, A., Pinder, A., et al. (2015). Aquatic communities in arid landscapes: local conditions, dispersal-traits and landscape configuration determine local biodiversity. Diversity and Distributions, 21, 12301241.Google Scholar
Noel, D. S., Martin, C. W., and Federer, C. A. (1986). Effects of forest clearcutting in New England on stream macroinvertebrates and periphyton. Environmental Management, 10, 661670.Google Scholar
Rooney, N. and McCann, K. S. (2012). Integrating food web diversity, structure and stability. Trends in Ecology and Evolution, 27, 4046.Google Scholar
Stark, J. D. (1985). A macroinvertebrate community index of water quality for stony streams. Water & Soil Miscellaneous Publication 87. Wellington, New Zealand: National Water and Soil Conservation Authority.Google Scholar
Stark, J. D. (1993). Performance of the Macroinvertebrate Community Index: effects of sampling method, sample replication, water depth, current velocity, and substratum on index values. New Zealand Journal of Marine and Freshwater Research, 27, 463478.Google Scholar
Stark, J. D. and Maxted, J. R. (2007). A user guide for the Macroinvertebrate Community Index. Wellington, New Zealand: Ministry for the Environment.Google Scholar
Tavares-Cromar, A. F. and Williams, D. D. (1996). The importance of temporal resolution in food web analysis: evidence from a detritus-based stream. Ecological Monographs, 66, 91113.Google Scholar
Tetzlaff, D., Malcolm, I. A., and Soulsby, C. (2007). Influence of forestry, environmental change and climatic variability on the hydrology, hydrochemistry and residence times of upland catchments. Journal of Hydrology, 346, 93111.Google Scholar
Thompson, R. and Starzomski, B. M. (2007). What does biodiversity actually do? A review for managers and policy makers. Biodiversity and Conservation, 16, 13591378.Google Scholar
Thompson, R. M. and Townsend, C. R. (2000). Is resolution the solution? The effect of taxonomic resolution on the calculated properties of three stream food webs. Freshwater Biology, 44, 413422.Google Scholar
Thompson, R. M. and Townsend, C. R. (2003). Impacts on stream food webs of native and exotic forest: an intercontinental comparison. Ecology, 84, 145161.Google Scholar
Thompson, R. M. and Townsend, C. R. (2005). Energy availability, spatial heterogeneity and ecosystem size predict food-web structure in streams. Oikos, 108, 137148.Google Scholar
Thompson, R. M., Phillips, N. R., and Townsend, C. R. (2009). Biological consequences of clear-cut logging around streams: moderating effects of management. Forest Ecology and Management, 257, 931940.Google Scholar
Townsend, C., Doledec, S., and Scarsbrook, M. (1997). Species traits in relation to temporal and spatial heterogeneity in streams: a test of habitat templet theory. Freshwater Biology, 37, 367387.Google Scholar
Townsend, C. R., Thompson, R. M., and Mcintosh, A. R. (1998). Disturbance, resource supply, and food web architecture in streams. Ecology Letters, 1, 200209.Google Scholar
Trayler, K. M. and Davis, J. A. (1998). Forestry impacts and the vertical distribution of stream invertebrates in south-western Australia. Freshwater Biology, 40, 331342.Google Scholar
Walker, B., Holling, C. S., Carpenter, S. R., and Kinzig, A. (2004). Resilience, adaptability and transformability in social-ecological systems. Ecology and Society, 9, 5.Google Scholar
Wardle, D. A. (1995). Impacts of disturbance on detritus food webs in agro-ecosystems of contrasting tillage and weed management practices. Advances in Ecological Research, 26, 105185.Google Scholar
Warren, P. H. (1994). Making connections in food webs. Trends in Ecology and Evolution, 9, 136141.Google Scholar
Williams, R. J. and Martinez, N. D. (2004). Limits to trophic levels and omnivory in complex food webs: theory and data. American Naturalist, 163, 458468.Google Scholar
Woodward, G., Speirs, D. C., and Hildrew, A. G. (2005). Quantification and resolution of a complex, size-structured food web. Advances in Ecological Research, 36, 85135.Google Scholar
Yodzis, P. (1981). The stability of real ecosystems. Nature, 289, 674676.Google Scholar
Yoon, I., Williams, R., Levine, E., et al. (2004). Webs on the web (WOW): 3D visualization of ecological networks on the WWW for collaborative research and education. Proceedings of SPIE – The International Society for Optical Engineering, 5295, 124132.Google Scholar
Zhang, Y., Richardson, J. S., and Pinto, X. (2009). Catchment-scale effects of forestry practices on benthic invertebrate communities in Pacific coastal streams. Journal of Applied Ecology, 46, 12921303.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
×