Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-75dct Total loading time: 0 Render date: 2024-05-07T20:12:53.582Z Has data issue: false hasContentIssue false

2 - The spatio-temporal dynamics of trophic control in large marine ecosystems

from Part II - Ecosystems

Published online by Cambridge University Press:  05 May 2015

Kenneth T. Frank
Affiliation:
Bedford Institute of Oceanography
Jonathan A. D. Fisher
Affiliation:
Institute of Memorial University of Newfoundland
William C. Leggett
Affiliation:
Queen’s University
Torrance C. Hanley
Affiliation:
Northeastern University, Boston
Kimberly J. La Pierre
Affiliation:
University of California, Berkeley
Get access

Summary

Introduction

The ways in which productivity, stability, population interactions, and community structure are regulated in ecosystems have been a central focus of ecology for over a century. At large spatial scales, major insights into these dynamics have been principally derived from analyses of changes induced from hunting, harvesting, and agricultural practices – so-called “natural experiments.” In terrestrial ecosystems estimates of the fraction of land transformed or degraded by human activity fall within the range of 39 to 75% (Vitousek et al., 1997; Ellis et al., 2010). Equally profound is the reality that up to 75% of the global oceans and in particular the continental shelf, transitional slope water areas, and reef habitats have been strongly impacted by human activity (Halpern et al., 2008).

One of the most widely studied human impacts has been the over-exploitation of large-bodied species. Berger et al. (2001) estimated that the spatial distribution of large mammalian carnivores that once played a dominant role in terrestrial ecosystems has declined by 95–99%. In the global oceans large predatory fish biomass may be as low as 10% of pre-industrial levels (Myers and Worm, 2003). These changes have created a vertical compaction and blunting of the trophic pyramid (Duffy, 2003; Chapter 14, this volume). On a global scale, these losses are attributable to a positive association between body size and sensitivity to population declines experienced by larger species which exhibit a greater susceptibility to decline or collapse as a consequence of their lower population densities, greater times to maturity, lower clutch sizes, and larger home ranges (Schipper et al., 2008). This reduction in the abundance of apex predators has led to abnormally high densities of their former prey in a wide range of ecosystems, which has, in turn, resulted in sometimes catastrophic changes in the ecosystems occupied. This has led some to conclude that large-bodied species are essential to the maintenance of ecosystem structure and stability (Hildrew et al., 2007; Estes et al., 2011).

Type
Chapter
Information
Trophic Ecology
Bottom-up and Top-down Interactions across Aquatic and Terrestrial Systems
, pp. 31 - 54
Publisher: Cambridge University Press
Print publication year: 2015

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

Anderson, S. C., Lotze, H. K. and Shackell, N. L. (2008). Evaluating the knowledge base for expanding low-trophic-level fisheries in Atlantic Canada. Canadian Journal of Fisheries and Aquatic Sciences, 65, 2553–2571.CrossRefGoogle Scholar
Ayón, P., Purca, S. and Guevara-Carrasco, R. (2004). Zooplankton volume trends off Peru between 1964 and 2001. ICES Journal of Marine Science, 61, 478–484.CrossRefGoogle Scholar
Bailey, D. M., Ruhl, H. A. and Snith, K. L. (2006). Long-term change in benthopelagic fish abundance in the abyssal northeast Pacific Ocean. Ecology, 87(3), 549–555.CrossRefGoogle ScholarPubMed
Bakun, A., Babcock, A. and Santora, C. (2009). Regulating a complex adaptive system via its wasp-waist: grappling with ecosystem-based management of the New England herring fishery. ICES Journal of Marine Science, 66, 1768–1775.CrossRefGoogle Scholar
Barange, M., Field, J. G., Harris, R. P., Hofmann, E. E. and Perry, R. I. (eds.) (2010). Marine Ecosystems and Global Change. Oxford: Oxford University Press.CrossRefGoogle Scholar
Barton, A. D., Pershing, A. J., Litchman, E., et al. (2013). The biogeography of marine plankton traits. Ecology Letters, 16, 522–534.CrossRefGoogle ScholarPubMed
Baum, J. K. and Worm, B. (2009). Cascading top-down effects of changing oceanic predator abundances. Journal of Animal Ecology, 78, 699–714.CrossRefGoogle ScholarPubMed
Bax, N. J. (1998). The significance and prediction of predation in marine fisheries. ICES Journal of Marine Science, 55, 997–1030.CrossRefGoogle Scholar
Benoît, H. P. and Swain, D. P. (2008). Impacts of environmental change and direct and indirect harvesting effects on the dynamics of a marine fish community. Canadian Journal of Fisheries and Aquatic Sciences, 65, 2088–2104.CrossRefGoogle Scholar
Berger, J., Stacey, P. B., Bellis, L. and Johnson, M. P. (2001). A mammalian predator-prey imbalance: grizzly bear and wolf extinction affect avian neotropical migrants. Ecological Applications, 11, 947–960.Google Scholar
Brodeur, R. D., Suchman, C. L., Reese, D. C., Miller, T. W. and Daly, E. A. (2008). Spatial overlap and trophic interactions between pelagic fish and large jellyfish in the northern California Current. Marine Biology, 154, 649–659.CrossRefGoogle Scholar
Brooks, J. L. and Dodson, S. I. (1965). Predation, body size, and composition of plankton. Science, 150, 28–35.CrossRefGoogle ScholarPubMed
Bundy, A. (2005). Structure and functioning of the eastern Scotian Shelf ecosystem before and after the collapse of groundfish stocks in the early 1990s. Canadian Journal of Fisheries and Aquatic Sciences, 62(7), 1453–1473.CrossRefGoogle Scholar
Caddy, J. F. and Garibaldi, L. (2000). Apparent changes in the trophic composition of world marine harvests: the perspective from the FAO capture database. Ocean & Coastal Management, 43, 615–655.CrossRefGoogle Scholar
Carpenter, S. R., Kitchell, J. F. and Hodgson, J. R. (1985). Cascading trophic interactions and lake productivity. Bioscience, 35, 634–649.CrossRefGoogle Scholar
Carr, M. H., Neigel, J. E., Estes, J. A., et al. (2003). Comparing marine and terrestrial ecosystems: implications for the design of coastal marine reserves. Ecological Applications, 13, S90–S107.CrossRefGoogle Scholar
Casini, M., Hjelm, J., Molinero, J.-C., et al. (2009). Trophic cascades promote threshold-like shifts in pelagic marine ecosystems. Proceedings of the National Academy of Sciences of the USA, 106, 197–202.CrossRefGoogle ScholarPubMed
Chassot, E., Mélin, F., Le Pape, O. and Gascuel, D. (2007). Bottom-up control regulates fisheries production at the scale of eco-regions in European seas. Marine Ecology Progress Series, 343, 45–55.CrossRefGoogle Scholar
Choi, J. S., Frank, K. T., Leggett, W. C. and Drinkwater, K. (2004). Transition to an alternate state in a continental shelf ecosystem. Canadian Journal of Fisheries and Aquatic Sciences, 61, 505–510.CrossRefGoogle Scholar
Christensen, V., Guenette, S., Heymans, J. J., et al. (2003). Hundred-year decline of North Atlantic predatory fishes. Fish and Fisheries, 4, 1–24.CrossRefGoogle Scholar
Claireaux, G., Webber, D. M., Lagardère, J.-P. and Kerr, S. R. (2000). Influence of water temperature and oxygenation on the aerobic metabolic scope of Atlantic cod (Gadus morhua). Journal of Sea Research, 44, 257–265.CrossRefGoogle Scholar
Cury, P. M., Shin, Y.-J., Planque, B., et al. (2008). Ecosystem oceanography for global change in fisheries. Trends in Ecology and Evolution, 23(6), 338–346.CrossRefGoogle ScholarPubMed
Daan, N., Gislason, H., Pope, J. G. and Rice, J. C. (2005). Changes in the North Sea fish community: evidence of indirect effects of fishing? ICES Journal of Marine Science, 62, 177–188.CrossRefGoogle Scholar
Darimont, C. T., Carlson, S. M., Kinnison, M. T., et al. (2009). Human predators outpace other agents of trait change in the wild. Proceedings of the National Academy of Sciences of the USA, 106, 952–954.CrossRefGoogle ScholarPubMed
Daskalov, G. M., Grishin, A. N., Rodionov, S. and Mihneva, V. (2007). Trophic cascades triggered by overfishing reveal possible mechanisms of ecosystem regime shifts. Proceedings of the National Academy of Sciences of the USA, 104, 10518–10523.CrossRefGoogle ScholarPubMed
deYoung, B., Heath, M., Werner, F., et al. (2004). Challenges of modelling ocean basin ecosystems. Science, 304, 1463–1466.CrossRefGoogle Scholar
Doney, S. C., Ruckelshaus, M., Duffy, J. E., et al. (2012). Climate change impacts on marine ecosystems. Annual Reviews of Marine Science, 4, 11–37.CrossRefGoogle ScholarPubMed
Drinkwater, K. F. (2006). The regime shift of the 1920s and 1930s in the North Atlantic. Progress in Oceanography, 68, 134–151.CrossRefGoogle Scholar
Duarte, C. M., Holmer, M., Olsen, Y., et al. (2009). Will the oceans help feed humanity? BioScience, 59, 967–976.CrossRefGoogle Scholar
Duffy, J. E. (2002). Biodiversity and ecosystem function: the consumer connection. Oikos, 99, 201–219.CrossRefGoogle Scholar
Duffy, J. E. (2003). Biodiversity loss, trophic skew and ecosystem functioning. Ecology Letters, 6, 680–687.CrossRefGoogle Scholar
Eliasen, K., Reinert, J., Gaard, E., et al. (2011). Sandeel as a link between primary production and higher trophic levels on the Faroe shelf. Marine Ecology Progress Series, 438, 185–194.CrossRefGoogle Scholar
Ellis, E. C., Goldewijk, K. K., Siebert, S., Lightman, D. and Ramankutty, N. (2010). Anthropogenic transformation of the biomes, 1700 to 2000. Global Ecology and Biogeography, 19, 589–606.Google Scholar
Elton, C. (1927). Animal Ecology. London: Sidwick and Jackson.Google Scholar
Eriksson, B. K., Lunggren, L., Sandstrom, A., et al. (2009). Declines in predatory fish promote bloom-forming macroalgae. Ecological Applications, 19, 1975–1988.CrossRefGoogle ScholarPubMed
Estes, J. A., Terborgh, J., Brashares, J. S., et al. (2011). Trophic downgrading of planet Earth. Science, 333, 301–306.CrossRefGoogle ScholarPubMed
Fauchald, P. (2010). Predator-prey reversal: a possible mechanism for ecosystem hysteresis in the North Sea. Ecology, 91, 2191–2197.CrossRefGoogle ScholarPubMed
Finenko, Z. Z., Piontkovski, S. A., Williams, R. and Mishonov, A. V. (2003). Variability of phytoplankton and mesozooplankton biomass in the subtropical and tropical Atlantic Ocean. Marine Ecology Progress Series, 250, 125–144.CrossRefGoogle Scholar
Finkel, Z. V., Beardall, J., Flynn, K. J., et al. (2010). Phytoplankton in a changing world: cell size and elemental stoichiometry. Journal of Plankton Research, 32, 119–137.CrossRefGoogle Scholar
Finni, T., Kononen, K., Olsonen, R. and Wallström, K. (2001). The history of cyanobacterial blooms in the Baltic Sea. Ambio 30, 172–178.CrossRefGoogle ScholarPubMed
Fisher, J. A. D., Frank, K. T., Petrie, B., Leggett, W. C. and Shackell, N. L. (2008). Temporal dynamics within a contemporary latitudinal diversity gradient. Ecology Letters, 11, 883–897.CrossRefGoogle ScholarPubMed
Fisher, J. A. D., Frank, K. T. and Leggett, W. C. (2010a). Breaking Bergmann's rule: truncation of Northwest Atlantic marine fish body sizes. Ecology, 91, 2499–2505.CrossRefGoogle ScholarPubMed
Fisher, J. A. D., Frank, K. T. and Leggett, W. C. (2010b). Global variation in marine fish body size and its role in biodiversity-ecosystem functioning. Marine Ecology Progress Series, 405, 1–13.CrossRefGoogle Scholar
Fogarty, M. J. and Murawski, S. A. (1998). Large-scale disturbance and the structure of marine systems: fishery impacts on Georges Bank. Ecological Applications, 8: S6–S22.CrossRefGoogle Scholar
Frank, K. T., Petrie, B., Choi, J. S. and Leggett, W. C. (2005). Trophic cascades in a formerly cod-dominated ecosystem. Science, 308, 1621–1623.CrossRefGoogle Scholar
Frank, K. T., Petrie, B., Shackell, N. L. and Choi, J. S. (2006). Reconciling differences in trophic control in mid-latitude marine ecosystems. Ecology Letters, 9, 1096–1105.CrossRefGoogle ScholarPubMed
Frank, K. T., Petrie, B. and Shackell, N. L. (2007). The ups and downs of trophic control in continental shelf ecosystems. Trends in Ecology and Evolution, 22(5), 236–242.CrossRefGoogle ScholarPubMed
Frank, K. T., Petrie, B., Fisher, J. A. D. and Leggett, W. C. (2011). Transient dynamics of an altered large marine ecosystem. Nature, 477, 86–89.CrossRefGoogle ScholarPubMed
Franks, P. J. (2002). NPZ models of plankton dynamics: their construction, coupling to physics, and application. Journal of Oceanography, 58, 379–387.CrossRefGoogle Scholar
Frederiksen, M., Edwards, M., Richardson, A. J., Halliday, N. C. and Wanless, S. (2006). From plankton to top predators: bottom-up control of a marine food web across four trophic levels. Journal of Animal Ecology, 75, 1259–1268.CrossRefGoogle ScholarPubMed
Fréon, P., Barange, M.Arístegui, J. and McIntyre, A. D. (2009). Eastern boundary upwelling ecosystems: integrative and comparative approaches. Progress in Oceanography, 83, 1–14.CrossRefGoogle Scholar
Gerber, L. R., Morissette, L. and Pauly, D. (2009). Should whales be culled to increase fishery yield? Science, 323, 880–881.CrossRefGoogle ScholarPubMed
Gislason, H., Sinclair, M., Sainsbury, K. and O'Boyle, R. N. (2000). Symposium overview: incorporating ecosystem objectives within fisheries management. ICES Journal of Marine Science, 57, 468–475.CrossRefGoogle Scholar
Gonzalez, A. and Loreau, M. (2009). The causes and consequences of compensatory dynamics in ecological communities. Annual Review of Ecology Evolution and Systematics, 40, 393–414.CrossRefGoogle Scholar
Hairston, N. G., Smith, F. E. and Slobodkin, L. B. (1960). Community structure, population control, and competition. American Naturalist, 94, 421–425.CrossRefGoogle Scholar
Halpern, B. S., Walbridge, S., Selkoe, K. A., et al. (2008). A global map of human impact on marine ecosystems. Science, 319, 948–952.CrossRefGoogle ScholarPubMed
Head, E. J. H. and Sameoto, D. D. (2007). Inter-decadal variability in zooplankton and phytoplankton abundance on the Newfoundland and Scotian shelves. Deep-Sea Research II, 57, 2686–2701.Google Scholar
Heath, M. R., Speirs, D. C. and Steele, J. H. (2013). Understanding patterns and processes in models of trophic cascades. Ecology Letters doi:10.1111/ele.12200Google ScholarPubMed
Heck, K. L. Jr. and Valentine, J. F. (2007). The primacy of top-down effects in shallow benthic ecosystems. Estuaries and Coasts, 30, 371–381.CrossRefGoogle Scholar
Hensen, V. (1887). Ueber die Bestimmung des Plankton's oder des im Meere treibenden Materials an Pflanzen und Tieren. Kommission zur wiss. Untersuchung der deutschen Meere, in Kiel, 1882–1886, Bericht 5, Vols. 12–16, pp. 1–107. Schmidt and Klaunig.Google Scholar
Hilborn, R. and Walters, C. J. (1992). Quantitative Fisheries Stock Assessment: Choice, Dynamics and Uncertainty. Norwell, MA: Kluwer Academic Publishers.CrossRefGoogle Scholar
Hildrew, A., Raffaelli, D. and Edmonds-Brown, R. (2007). Body Size: The Structure and Function of Aquatic Ecosystems. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Johannesen, E., Ingvaldsen, R. B., Bogstad, B., et al. (2012). Changes in Barents Sea ecosystem state, 1970–2009: climate fluctuations, human impact, and trophic interactions. ICES Journal of Marine Science, 69(5), 880–889.CrossRefGoogle Scholar
Johnson, N. A., Campbell, J. W., Moore, T. S., et al. (2007). The relationship between the standing stock of the deep-sea macrobenthos and surface production in the western North Atlantic. Deep-Sea Research I, 54, 1350–1360.CrossRefGoogle Scholar
Kemp, W. M., Boynton, W. R., Adolf, J. E., et al. (2005). Eutrophication of Chesapeake Bay: historical trends and ecological interactions. Maine Ecology Progress Series, 303, 1–29.Google Scholar
Knowlton, N. (2004). Multiple stable states and the conservation of marine ecosystems. Progress in Oceanography, 60, 387–396.CrossRefGoogle Scholar
Köster, F. W. and Möllmann, C. (2000). Trophodynamic control by clupeid predators on recruitment success in Baltic cod? ICES Journal of Marine Science, 57, 310–323.CrossRefGoogle Scholar
Laptikhovsky, V., Arkhipkin, A. and Brickle, P. (2013). From small bycatch to main commercial species: explosion of stocks of rock cod Patagonotothen ramsayi (Regan) in the southwest Atlantic. Fisheries Research, 147, 399–403.CrossRefGoogle Scholar
Legendre, L. and Michaud, J. (1998). Flux of biogenic carbon in oceans: size-dependent regulation by pelagic food webs. Marine Ecology Progress Series, 164, 1–11.CrossRefGoogle Scholar
Llope, M., Licandro, P., Chan, K.-S. and Stenseth, N. C. (2012). Spatial variability of the plankton trophic interaction in the North Sea: a new feature after the early 1970s. Global Change Biology, 18, 106–117.CrossRefGoogle Scholar
Lynman, C. P., Gibbons, M. J., Axelsen, B. E., et al. (2006). Jellyfish overtake fish in a heavily fished ecosystem. Current Biology, 16, R492–493.Google Scholar
Manning, F. (2012). The sustainable management of grey seal populations: a path toward the recovery of cod and other groundfish stocks. Report of the Standing Senate Committee on Fisheries and Oceans. 42 pp.
McCann, K. S., Rasmussen, J. B. and Umbanhowar, J. (2005). The dynamics of spatially coupled food webs. Ecology Letters, 8, 513–523.CrossRefGoogle ScholarPubMed
Micheli, F. (1999). Eutrophication, fisheries, and consumer-resource dynamics in marine pelagic ecosystem. Science, 285, 1396–1398.CrossRefGoogle Scholar
Micheli, F., Benedetti-Cecchi, L., Gambaccini, S., et al. (2005). Cascading human impacts, marine protected areas, and the structure of Mediterranean reef assemblages. Ecological Monographs, 75, 81–102.CrossRefGoogle Scholar
Minto, C. and Worm, B. (2012). Interactions between small pelagic fish and young cod across the North Atlantic. Ecology, 93, 2139–2154.CrossRefGoogle Scholar
Möllmann, C., Müller-Karulis, B., Kornilovs, G. and St. John, M. (2008). Effects of climate and overfishing on zooplankton dynamics and ecosystem structure: regime shift, trophic cascade, and feedback loops in a simple ecosystem. ICES Journal of Marine Scence, 65, 302–310.Google Scholar
Mora, C., Tittensor, D. P., Adl, S., Simpson, A. G. B. and Worm, B. (2011). How many species are there on Earth and in the ocean? PLoS Biology, 9(8), e1001127.CrossRefGoogle ScholarPubMed
Myers, R. A., Mertz, G. and Fowlow, P. S. (1997). Maximum population growth rates and recovery times for Atlantic cod, Gadus morhua. Fishery Bulletin, 95, 762–772.Google Scholar
Myers, R. A. and Worm, B. (2003). Rapid worldwide depletion of predatory fish communities. Nature, 423, 280–283.CrossRefGoogle ScholarPubMed
Myers, R. A. and Worm, B. (2005). Extinction, survival, or recovery of large predatory fish. Philosophical Transactions of the Royal Society, Series B, 360, 13–20.CrossRefGoogle Scholar
Nixon, S. W. (1988). Physical energy inputs and the comparative ecology of lake and marine ecosystems. Limnology and Oceanography, 33, 1005–1025.Google Scholar
Österblom, H., Casini, M., Olsson, O. and Bignert, A. (2006). Fish, seabirds and trophic cascades in the Baltic Sea. Marine Ecology Progress Series, 323, 233–238.CrossRefGoogle Scholar
Österblom, H., Olsson, O., Blenckner, T. and Furness, R. W. (2008). Junk-food in marine ecosystems. Oikos, 117, 967–977.CrossRefGoogle Scholar
Paine, R. T. (1980). Food webs: linkage, interaction strength and community infrastructure. Journal of Animal Ecology, 49, 667–685.CrossRefGoogle Scholar
Pauly, D., Christensen, V., Guenette, S., et al. (2002). Towards sustainability in world fisheries. Nature, 418, 689–695.CrossRefGoogle ScholarPubMed
Pershing, A. J., Head, E. H., Greene, C. H. and Jossi, J. W. (2010). Pattern and scale of variability among Northwest Atlantic shelf plankton communities. Journal of Plankton Research, 32, 1661–1674.CrossRefGoogle Scholar
Peterson, W. T. and Schwing, F. B. (2003). A new climate regime in northeast Pacific ecosystems. Geophysical Resarch Letters, 30(17), 1896.Google Scholar
Petrie, B., Frank, K. T., Shackell, N. L. and Leggett, W. C. (2009). Structure and stability in exploited marine ecosystems: quantifying critical transitions. Fisheries Oceanography, 18(2), 83–101.CrossRefGoogle Scholar
Polis, G. A., Sears, A. L. W., Huxel, G. R., Strong, D. R. and Maron, J. (2000). When is a trophic cascade a trophic cascade? Trends in Ecology and Evolution, 15(11), 473–475.CrossRefGoogle ScholarPubMed
Reid, P. C., Battle, E. J. V., Batten, S. D. and Brander, K. M. (2000). Impacts of fisheries on plankton community structure. ICES Journal of Marine Science, 57, 495–502.CrossRefGoogle Scholar
Richardson, A. J. and Schoeman, D. S. (2004). Climate impact on plankton ecosystems in the Northeast Atlantic. Science, 305, 1609–1612.CrossRefGoogle ScholarPubMed
Richardson, A. J., Bakun, A., Hays, G. C. and Gibbons, M. J. (2009). The jellyfish joyride: causes, consequences and management responses to a more gelatinous future. Trends in Ecology and Evolution, 24(6), 312–322.CrossRefGoogle ScholarPubMed
Ritchie, E. G. and Johnson, C. N. (2009). Predator interactions, mesopredator release and biodiversity conservation. Ecology Letters, 12, 982–998.CrossRefGoogle ScholarPubMed
Romanuk, T. N., Hayward, A. and Hutchings, J. A. (2011). Trophic level scales positively with body size in fishes. Global Ecology and Biogeography, 20, 231–240.CrossRefGoogle Scholar
Salomon, A. K., Gaichas, S. K., Shears, N. T., et al. (2010). Key features and context-dependence of fishery-induced trophic cascades. Conservation Biology, 24(2), 382–394.CrossRefGoogle ScholarPubMed
Savenkoff, C., Swain, D. P., Hanson, J. M., et al. (2007). Effects of fishing and predation in a heavily exploited ecosystem: comparing periods before and after the collapse of groundfish in the southern Gulf of St. Lawrence (Canada). Ecological Modeling, 204, 115–128.CrossRefGoogle Scholar
Schipper, J., Chanson, J. S., Chiozza, F., et al. (2008). The status of the world's land and marine mammals: diversity, threat, and knowledge. Science, 322, 225–230.CrossRefGoogle ScholarPubMed
Shackell, N. L. and Frank, K. T. (2007). Compensation in exploited marine fish communities on the Scotian Shelf, Canada. Marine Ecology Progress Series, 336, 235–247.CrossRefGoogle Scholar
Shackell, N. L., Frank, K. T., Fisher, J. A. D., Petrie, B. and Leggett, W. C. (2010). Decline in top predator body size and changing climate alter trophic structure in an oceanic ecosystem. Proceedings of the Royal Society, Series B, 277, 1353–1360.CrossRefGoogle Scholar
Shears, N. T., Babcock, R. C. and Salomon, A. K. (2008). Context-dependent effects of fishing: variation in trophic cascades across environmental gradients. Ecological Applications, 18, 1860–1873.CrossRefGoogle ScholarPubMed
Shurin, J. B., Borer, E. T., Seabloom, E. W., et al. (2002). A cross-ecosystem comparison of the strength of trophic cascades. Ecology Letters, 5, 785–791.CrossRefGoogle Scholar
Sinclair, A. R. E. and Krebs, C. J. (2002). Complex numerical responses to top-down and bottom-up processes in vertebrate populations. Philosophical Transactions of the Royal Society of London Series B, 357, 1221–1231.CrossRefGoogle ScholarPubMed
Steele, J. H. (1998). Regime shifts in marine ecosystems. Ecological Applications, 8, S33–S36.CrossRefGoogle Scholar
Strong, D. R. (1992). Are trophic cascades all wet? Differentiation and donor-control in speciose ecosystems. Ecology, 73, 747–754.CrossRefGoogle Scholar
Strong, D. R. and Frank, K. T. (2010). Human involvement in food webs. Annual Review of Environmental Resources, 35, 1–23.CrossRefGoogle Scholar
Swain, D. P. and Mohn, R. K. (2012). Forage fish and the factors governing recovery of Atlantic cod (Gadus morhua) on the eastern Scotian Shelf. Canadian Journal of Fisheries and Aquatic Sciences, 69, 997–1001.CrossRefGoogle Scholar
Swain, D. P. and Sinclair, A. F. (2000). Pelagic fishes and the cod recruitment dilemma in the Northwest Atlantic. Canadian Journal of Fisheries and Aquatic Sciences, 57, 1321–1325.CrossRefGoogle Scholar
Tremblay-Boyer, L., Gascuel, D., Watson, R., Christensen, V. and Pauly, D. (2011). Modelling the effects of fishing on the biomass of the world's oceans from 1950 to 2006. Marine Ecology Progress Series, 442, 169–185.CrossRefGoogle Scholar
Turner, J. T. and Granéli, E. (2006). “Top-down” predation control on marine harmful algae. In Ecology of Harmful Algae. Ecological Studies, Vol. 189. Berlin Heidelberg: Springer-Verlag, pp. 355–366.Google Scholar
Verity, P. G. (1998). Why is relating plankton community structure to pelagic production so problematic? South African Journal of Marine Science, 19, 333–338.CrossRefGoogle Scholar
Verity, P. G. and Smetacek, V. (1996). Organism life cycles, predation, and the structure of marine pelagic ecosystems. Marine Ecology Progress Series, 130, 277–293.CrossRefGoogle Scholar
Vitousek, P. M., Mooney, H. A., Lubchenco, J. and Melillo, J. M. (1997). Human domination of earth's ecosystems. Science, 277, 494–499.CrossRefGoogle Scholar
Walters, C. and Kitchell, J. F. (2001). Cultivation/depensation effects on juvenile survival and recruitment: implications for the theory of fishing. Canadian Journal of Fisheries and Aquatic Sciences, 58, 39–50.CrossRefGoogle Scholar
Ware, D. M. and Thompson, R. E. (2005). Bottom-up ecosystem trophic dynamics determine fish production in the Northeast Pacific. Science, 308, 1280–1284.CrossRefGoogle ScholarPubMed
Webb, T. J., Berghe, E. V. and O'Dor, R. (2010). Biodiversity's big wet secret: the global distribution of marine biological records reveals chronic under-exploration of the deep pelagic ocean. PLoS One, 5(8), e10223.CrossRefGoogle ScholarPubMed
White, T. C. R. (1978). The importance of a relative shortage of food in animal ecology. Oecologia, 33, 71–86.CrossRefGoogle ScholarPubMed
Worm, B. and Myers, R. A. (2003). Meta-analysis of cod-shrimp interactions reveals top-down control in oceanic food webs. Ecology, 84, 162–173.CrossRefGoogle Scholar
Yodzis, P. (2001). Must top predators be culled for the sake of fisheries?Trends in Ecology and Evolution, 16, 78–84.CrossRefGoogle ScholarPubMed

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
×