Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-9pm4c Total loading time: 0 Render date: 2024-04-26T19:30:40.532Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 June 2012

Marc Mangel
Affiliation:
University of California, Santa Cruz
Get access
Type
Chapter
Information
The Theoretical Biologist's Toolbox
Quantitative Methods for Ecology and Evolutionary Biology
, pp. 323 - 368
Publisher: Cambridge University Press
Print publication year: 2006

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

Abrahams, M. V. and Healey, M. C. (1993). Some consequences of variation in vessel density: a manipulative field experiment. Fisheries Research, 15, 315–322.CrossRefGoogle Scholar
Abramowitz, M. and Stegun, I. (1974). Handbook of Mathematical Functions. New York: Dover Publications.Google Scholar
Acosta, C. A. (2002). Spatially explicit dispersal dynamics and equilibrium population sizes in marine harvest refuges. ICES Journal of Marine Science, 59, 458–468.CrossRefGoogle Scholar
Ade, R. (1989). The Trout and Salmon Handbook. New York: Facts on File.Google Scholar
Adkison, M. D. and Zhenming, S. (2001). A comparison of salmon escapement estimates using a hierarchical Bayesian approach versus separate maximum likelihood estimation of each year's return. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1663–1671.CrossRefGoogle Scholar
Adler, F. R. (1993). Migration alone can produce persistence of host–parasitoid models. American Naturalist, 141, 642–650.CrossRefGoogle Scholar
Ahmed, A. M., Baggot, S. L., Maingon, R. and Hurd, H. (2002). The costs of mounting an immune response are reflected in the reproductive fitness of the mosquito Anopheles gambiae. Oikos, 97, 371–377.CrossRefGoogle Scholar
Allen, J. E. and Maizels, R. M. (1997). Th1-Th2: reliable paradigm or dangerous dogma?Immunology Today, 18, 387–392.CrossRefGoogle ScholarPubMed
Alpatov, W. W. and Pearl, R. (1929). Experimental studies on the duration of life. XII. Influence of temperature during the larval period and adult life on the duration of the life of the imago of drosophila melanogaster. American Naturalist, 63, 37–67.CrossRefGoogle Scholar
Anand, P. (2002). Decision-making when science is ambiguous. Science, 295, 1839.CrossRefGoogle ScholarPubMed
Anderson, J. J. (1992). A vitality-based stochastic model for organism survival. In DeAngelis, D. L. and Gross, L. J. (editors) Individual-Based Models and Approaches in Ecology. New York: Chapman and Hall, pp. 256–277.CrossRefGoogle Scholar
Anderson, J. J. (2000). A vitality-based model relating stressors and environmental properties to organismal survival. Ecological Monographs, 70, 445–470.CrossRefGoogle Scholar
Anderson, L. G. (1991a). Efficient policies to maintain total allowable catches in ITQ fisheries with at-sea processing. Land Economics, 67, 141–157.CrossRefGoogle Scholar
Anderson, L. G. (1991b). A note on market power in ITQ fisheries. Journal of Environmental Economics and Management, 21, 291–296.CrossRefGoogle Scholar
Anderson, R. M. (editor) (1982). Population Dynamics of Disease: Theory and Applications. London: Chapman and Hall.CrossRefGoogle Scholar
Anderson, R. M. (1991). The Kermack–McKendrick epidemic threshold theorem. Bulletin of Mathematical Biology, 53, 3–32.CrossRefGoogle ScholarPubMed
Anderson, R. M. and May, R. M. (1978). Regulation and stability of host–parasite population interactions. I Regulatory processes. Journal of Animal Ecology, 47, 219–247.CrossRefGoogle Scholar
Anderson, R. M. and May, R. M. (1991). Infectious Diseases of Humans. Dynamics and Control. Oxford: Oxford University Press.Google Scholar
Anonymous, (2002). The State of the World Fisheries and Aquaculture. Rome: Food and Agriculture Organization of the United Nations.
Anscombe, F. J. (1950). Sampling theory of the negative binomial and logarithmic series distributions. Biometrika, 37, 358–382.CrossRefGoogle ScholarPubMed
Antia, R. and Lipsitch, M. (1997). Mathematical models of parasite responses to host immune defenses. Parasitology, 115, S115–S167.CrossRefGoogle Scholar
Antia, R., Levin, B. R. and May, R. M. (1994). Within-host population dynamics and the evolution and maintenance of microparasite virulence. American Naturalist, 144, 457–472.CrossRefGoogle Scholar
Antia, R., Nowak, M. A. and Anderson, R. M. (1996). Antigenic variation and the within-host dynamics of parasites. Proceedings of the National Academy of Sciences, 93, 985–989.CrossRefGoogle ScholarPubMed
Apollonio, S. (1994). The use of ecosystem characteristics in fisheries management. Reviews in Fisheries Science, 2, 157–180.CrossRefGoogle Scholar
Apostolaki, P., Milner-Gulland, E. J., McAllister, M. K. and Kirkwood, G. P. (2002). Modelling the effects of establishing a marine reserve for mobile species. Canadian Journal of Fisheries and Aquatic Sciences, 59, 405–415.CrossRefGoogle Scholar
Arnold, L. (1973). Stochastic Differential Equations. New York: John Wiley and Sons.Google Scholar
Arnold, L. (1974). Stochastic Differential Equations: Theory and Applications. London: Wiley Interscience.Google Scholar
Austin, D. J., White, N. J. and Anderson, R. M. (1998). The dynamics of drug action on the within-host population growth of infectious agents: melding pharmacokinetics with pathogen population dynamics. Journal of Theoretical Biology, 194, 313–339.CrossRefGoogle ScholarPubMed
Babcock, E. A. and Pikitch, E. K. (2000). A dynamic programming model of fishing strategy choice in a multispecies trawl fishery with trip limits. Canadian Journal of Fisheries and Aquatic Sciences, 57, 357–370.CrossRefGoogle Scholar
Bailey, N. T. J. (1953). The total size of a general stochastic epidemic. Biometrika, 40, 177–185.CrossRefGoogle Scholar
Bakun, A. (1996). Patterns in the Ocean. Oceanic Process and Marine Population Dynamics. La Jolla, CA: California Sea Grant.Google Scholar
Barlow, N. D. (1993). A model for the spread of bovine Tb in New Zealand possum populations. Journal of Applied Ecology, 30, 156–164.CrossRefGoogle Scholar
Barlow, N. D. (2000). Non-linear transmission and simple models for bovine tuberculosis. Journal of Animal Ecology, 69, 703–713.CrossRefGoogle Scholar
Barnard, G. A. and Bayes, T. (1958). Studies in the history of probability and statistics. IX. Thomas Bayes's essay towards solving a problem in the doctrine of chances. Biometrika, 45, 293–315.CrossRefGoogle Scholar
Barrowman, N. J. and Myers, R. A. (2000). Still more spawner-recruitment curves: the hockey stick and its generalizations. Canadian Journal of Fisheries and Aquatic Sciences, 57, 665–676.CrossRefGoogle Scholar
Bartlett, B. R. (1964). Patterns in the host-feeding habit of adult parasite hymenoptera. Annals of the Entomological Society of America, 57, 344–350.CrossRefGoogle Scholar
Bartlett, M. S. (1957). Measles periodicity and community size. Journal of the Royal Statistical Society A, 120, 48–70.CrossRefGoogle Scholar
Basar, T. and Olsder, G. J. (1982). Dynamic Noncooperative Game Theory. New York: Academic Press.Google Scholar
Bazykin, A. D. (1998). Nonlinear Dynamics of Interacting Populations. Singapore and River Edge, NJ: World Scientific.Google Scholar
Beamish, R. J. (2002). Obituary. William Edwin Ricker OC, FRSC, LLD, DSc. Journal of Fish Biology, 60, 285–287.CrossRefGoogle Scholar
Begon, M., Bowers, R. G., Kadianakis, N. and Hodgkinson, D. E. (1992). Disease and community structure: the importance of host self-regulation in a host–host pathogen model. The American Naturalist, 139, 1131–1150.CrossRefGoogle Scholar
Begon, M., Feore, S. M., Brown, K., et al. (1998). Population and transmission dynamics of cowpox in bank voles: testing fundamental assumptions. Ecology Letters, 1, 82–86.CrossRefGoogle Scholar
Begon, M., Hazel, S. M., Baxby, D., et al. (1999). Transmission dynamics of a zoonotic pathogen within and between wildlife host species. Proceedings of the Royal Society of London, B266, 1939–1945.CrossRefGoogle ScholarPubMed
Behnke, R. J. (1992). Native Trout of Western North America. Bethesda, MD: American Fisheries Society.Google Scholar
Behnke, R. J. (2002). Trout and Salmon of North America. New York: Free Press.Google Scholar
Belew, R. K. and Mitchell, M. (editors) (1996). Adaptive Individual in Evolving Populations: Models and Algorithms. Reading, MA: Addison Wesley.Google Scholar
Bellman, R. E. (1984). Eye of the Hurricane: An Autobiography. Singapore: World Scientific.CrossRefGoogle Scholar
Bellman, R. and Cooke, K. L. (1963). Differential-Difference Equations. New York: Academic Press.Google Scholar
Bellows, T. S. and Hassell, M. P. (1988). The dynamics of age-structured host-parasitoid interactions. Journal of Animal Ecology, 57, 259–268.CrossRefGoogle Scholar
Bender, C. M. and Orszag, S. A. (1978). Advanced Mathematical Methods for Scientists and Engineers. New York: McGraw Hill.Google Scholar
Berger, R., Quack, M. and Tschumper, G. S. (2000). Electroweak quantum chemistry for possible precursor molecules in the evolution of biomolecular homochirality. Helvetica Chimica Acta, 83, 1919–1950.3.0.CO;2-D>CrossRefGoogle Scholar
Bernoulli, D. (1738/1954). Exposition of a new theory on the measurement of risk. Economietrica, 22, 23–36.CrossRefGoogle Scholar
Bertsekas, D. P (1976). Dynamic Programming and Stochastic Control. New York: Academic Press.Google Scholar
Bertsekas, D. P. (1995). Dynamic Programming and Optimal Control. Volumes One and Two. Belmont, MA: Athena Scientific.Google Scholar
Beverton, R. J. H. (1992). Patterns of reproductive strategy parameters in some marine teleost fishes. Journal of Fish Biology, 41 (Supplement B), 137–160.CrossRefGoogle Scholar
Beverton, R. J. H. (1994). Notes on the Use of Theoretical Models in the Study of the Dynamics of Exploited Fish Populations. From the Lectures of R. J. H. Beverton presented at U.S. Fishery Laboratory, Beaufort, North Carolina, Bureau of Commercial Fisheries 1951. Beaufort, NC: Marine Fisheries Section, American Fisheries Society.Google Scholar
Beverton, R. J. H. and Holt, S. J. (1957). On the Dynamics of Exploited Fish Populations. London: HMSO.Google Scholar
Beverton, R. J. H. and Holt, S. J. (1959). A review of the lifespans and mortality rates of fish in nature, and their relation to growth and other physiological characteristics. In Wolstenholme, G. E. W. and O'Connor, M. (editors) CIBA Foundation Colloquia on Ageing. London: CIBA Foundation, pp. 142–174.Google Scholar
Bharucha-Reid, A. T. (1997 (1960)). Elements of Stochastic Processes. New York: Dover.Google Scholar
Bjorksted, E. P. (2000). Stock–recruitment relationships for life cycles that exhibit concurrent density dependence. Canadian Journal of Fisheries and Aquatic Sciences, 57, 459–467.CrossRefGoogle Scholar
Bjørnstad, O. N. and Grenfell, B. T. (2001). Noisy clockwork: time series analysis of population fluctuations in animals. Science, 293, 638–643.CrossRefGoogle ScholarPubMed
Black, F. and Scholes, M. (1972). The pricing of options and corporate liabilities. The Journal of Political Economy, 81, 637–654.CrossRefGoogle Scholar
Blackmond, D. G., McMillan, C. R., Ramdeechul, S., Schorm, A. and Brown, J. M. (2001). Origins of asymmetric amplifications in autocatalytic alkylzinc additions. Journal of the American Chemical Society, 123, 10 103–10 104.CrossRefGoogle Scholar
Bleistein, N. and Handelsman, R. A. (1975). Asymptotic Expansions of Integrals. New York: Holt, Rinehart and Winston.Google Scholar
Blythe, S. P., Nisbet, R. M. and Gurney, W. S. C. (1984). The dynamics of population models with distributed maturation periods. Theoretical Population Biology, 25, 289–311.CrossRefGoogle ScholarPubMed
Bonsall, M. P. and Hastings, A. (2004). Demographic and environmental stochasticity in predator–prey metapopulation dynamics. Journal of Animal Ecology, 73, 1043–1055.CrossRefGoogle Scholar
Bonsall, M. B., Hassell, M. P. and Asefa, G. (2002). Ecological trade-offs, resource partitioning, and coexistence in a host–parasitoid assemblage. Ecology, 83, 925–934.Google Scholar
Boreman, J., Nakashima, B. S., Wilson, J. A. and Kendall, R. L. (editors) (1997). Northwest Atlantic Groundfish: Perspectives on a Fishery Collapse. Bethesda, MD: American Fisheries Society.Google Scholar
Botsford, L. W., Hastings, A. and Gaines, S. D. (2001). Dependence of sustainability on the configuration of marine reserves and larval dispersal distance. Ecology Letters, 4, 144–150.CrossRefGoogle Scholar
Bouloux, C., Langlais, M. and Silan, P. (1998). A marine host-parasite model with direct biological cycle and age structure. Ecological Modelling, 107, 73–86.CrossRefGoogle Scholar
Bradford, M., Myers, R. A. and Irvine, J. R. (2000). Reference points for coho salmon (Oncorhynchus kisutch) harvest rates and escapement goals based on freshwater production. Canadian Journal of Fisheries and Aquatic Sciences, 57, 677–686.CrossRefGoogle Scholar
Brenner, S. (1999). Theoretical biology in the third millennium. Philosophical Transactions of the Royal Society of London, Series B, Biological Sciences, 354, 1963–1965.CrossRefGoogle ScholarPubMed
Briggs, C. J. (1993). Competition among parasitoid species on a stage-structure host and its effect on host suppression. American Naturalist, 141, 372–397.CrossRefGoogle Scholar
Briggs, C. J. and Hoopes, M. F. (2004). Stabilizing effects in spatial parasitoid-host and predator-prey models: a review. Theoretical Population Biology, 65, 299–315.CrossRefGoogle ScholarPubMed
Brinkman, H. C. (1956). Brownian motion in a field of force and the diffusion theory of chemical reactions, II. Physica, 22, 149–155.CrossRefGoogle Scholar
Brodziak, J. K. T., Overholtz, W. J. and Rago, P. J. (2001). Does spawning stock affect recruitment of New England groundfish. Canadian Journal of Fisheries and Aquatic Sciences, 58, 306–318.CrossRefGoogle Scholar
Brookhart, M. A., Hubbard, A. E., Laan, M. J., Colford, J. M. Jr. and Eisenberg, J. N. S. (2002). Statistical estimation of parameters in a disease transmission model: analysis of a Cryptosporidium outbreak. Statistics in Medicine, 21, 3627–3638.CrossRefGoogle Scholar
Brooks, E. N. (2002). Using reproductive values to define optimal harvesting for multisite density-dependent populations: example with a marine reserve. Canadian Journal of Fisheries and Aquatic Sciences, 59, 875–885.CrossRefGoogle Scholar
Brown, R. (1828). A brief account of microscopical observations made in the months of June, July, and August, 1827, on the particles contained in the pollen of plants; and on the general existence of active molecules in organic and inorganic bodies. The Philosophical Magazine, 4, 161–173.CrossRefGoogle Scholar
Brunet, L. R., Dunne, D. W. and Pearce, E. J. (1998). Cytokine interaction and immune response during Schistosoma mansoni infection. Parasitology Today, 14, 422–427.CrossRefGoogle ScholarPubMed
Bucklew, J. A. (1990). Large Deviation Techniques in Decision, Simulation, and Estimation. New York: Wiley Interscience.Google Scholar
Burnham, K. P. and Anderson, D. R. (1998). Model Selection and Inference. A Practical Information–Theoretic Approach. New York: Springer Verlag.CrossRefGoogle Scholar
Bynum, W. F. (2002). Mosquitoes bite more than once. Science, 295, 47.CrossRefGoogle ScholarPubMed
Caddy, J. F. (2002). Limit reference points, traffic lights, and holistic approaches to fisheries management with minimal stock assessment input. Fisheries Research, 56, 133–137.CrossRefGoogle Scholar
Calder, W. A. I. (1984). Size, Function, and Life History. Cambridge, MA: Harvard University Press.Google Scholar
Callaway, D. S. and Perelson, A. S. (2002). HIV-1 infection and low steady state viral loads. Bulletin of Mathematical Biology, 64, 29–64.CrossRefGoogle ScholarPubMed
Caraco, T., Duryea, M. C., Glavanakov, S., Maniatty, W. and Szymanski, B. K. (2001). Host spatial heterogeneity and the spread of vector-borne infection. Theoretical Population Biology, 59, 185–206.CrossRefGoogle ScholarPubMed
Cardinale, M. and Arrhenius, F. (2000). Decreasing weight-at-age of Atlantic herring (Clupea harengus) from the Baltic Sea between 1986 and 1996: a statistical analysis. ICES Journal of Marine Science, 57, 882–893.CrossRefGoogle Scholar
Cardinale, M. and Modin, J. (1999). Changes in size-at-maturity of Baltic cod (Gadus morhua) during a period of large variations in stock size and environmental conditions. Fisheries Research, 41, 285–295.CrossRefGoogle Scholar
Carey, J. R. (2001). Insect biodemography. Annual Review of Entomology, 46, 79–110.CrossRefGoogle ScholarPubMed
Carey, J. R. (2003). Longevity. The Biology and Demography of Life Span. Princeton, NJ: Princeton University Press.Google Scholar
Carey, J. R. and Judge, D. S. (2001). Principles of biodemography with special reference to human longevity. Population: An English Selection, 13, 9–40.Google Scholar
Carpenter, S. R., Ludwig, D. and Brock, W. A. (1999). Management of eutrophication for lakes subject to potentially irreversible change. Ecological Applications, 9, 751–771.CrossRefGoogle Scholar
Carslaw, H. S. and Jaeger, J. C. (1959). Conduction of Heat in Solids. Oxford: Oxford University Press.Google Scholar
Catteruccia, F., Godfray, C. and Crisanti, A. (2003). Impact of genetic manipulation of the fitness of Anopheles stephensi mosquitoes. Science, 299, 1225–1227.CrossRefGoogle ScholarPubMed
Caughley, G. (1994). Directions in conservation biology. Journal of Animal Ecology, 63, 215–244.CrossRefGoogle Scholar
Chapman, R. N. (1928). The quantitative analysis of environmental factors. Ecology, 9, 111–122.CrossRefGoogle Scholar
Charles, A. T. (1992). Fishery conflicts. A unified framework. Marine Policy, 16, 379–393.CrossRefGoogle Scholar
Charles, S., Morand, S., Chasse, J. L. and Auger, P. (2002). Host patch selection induced by parasitism: basic reproduction ratio R0 and optimal virulence. Theoretical Population Biology, 62, 97–109.CrossRefGoogle Scholar
Charlesworth, B. (1990). Optimization models, quantitative genetics, and mutation. Evolution, 44, 520–538.CrossRefGoogle ScholarPubMed
Charlesworth, B. (1994). Evolution in Age-Structured Populations. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Charnov, E. (1993). Life History Invariants. New York: Oxford University Press.Google Scholar
Charnov, E. L. (1976). Optimal foraging, the marginal value theorem. Theoretical Population Biology, 9, 129–136.CrossRefGoogle ScholarPubMed
Charnov, E. L. and Skinner, S. W. (1984). Evolution of host selection and clutch size in parasitoid wasps. Florida Entomologist, 67, 5–21.CrossRefGoogle Scholar
Charnov, E. L. and Skinner, S. W. (1985). Complementary approaches to the understanding of parasitoid oviposition decisions. Environmental Entomology, 14, 383–391.CrossRefGoogle Scholar
Charnov, E. L. and Skinner, S. W. (1988). Clutch size in parasitoids: the egg production rate as a constraint. Evolutionary Ecology, 2, 167–174.CrossRefGoogle Scholar
Chattopadhyay, J. and Sarkar, R. R. (2003). Chaos to order: preliminary experiments with a population dynamics models of three trophic levels. Ecological Modelling, 163, 45–50.CrossRefGoogle Scholar
Cheever, J. (1978). The Stories of John Cheever. New York: Alfred A. Knopf.Google Scholar
Chen, D. G., Irvine, J. R. and Cass, A. J. (2002). Incorporating Allee effects in fish stock–recruitment models and applications for determining reference points. Canadian Journal of Fisheries and Aquatic Sciences, 59, 242–249.CrossRefGoogle Scholar
Chen, Y., Jiao, Y. and Chen, L. (2003). Developing robust frequentist and Bayesian fish stock assessment methods. Fish and Fisheries, 4, 105–120.CrossRefGoogle Scholar
Christensen, V., Guénette, S., Heymans, J. J., et al. (2003). Hundred-year decline of North Atlantic predatory fishes. Fish and Fisheries, 4, 1–24.CrossRefGoogle Scholar
Clark, C. W. (1973). The economics of overexploitation. Science, 181, 630–634.CrossRefGoogle ScholarPubMed
Clark, C. W. (1985). Bioeconomic Modelling and Fisheries Management. New York: Wiley Interscience.Google Scholar
Clark, C. W. (1990). Mathematical Bioeconomics, 2nd edn. New York: Wiley Interscience.Google Scholar
Clark, C. W. (2006). The Worldwide Crisis in Fisheries. New York: Cambridge University Press.Google Scholar
Clark, C. W. and Mangel, M. (1979). Aggregation and fishery dynamics: a theoretical study of schooling and the purse seine tuna fisheries. Fishery Bulletin, 77, 317–337.Google Scholar
Clark, C. W. and Mangel, M. (2000). Dynamic State Variable Models in Ecology. Methods and Applications. New York: Oxford University Press.Google Scholar
Clark, W. G. (1991). Groundfish exploitation rates based on life history parameters. Canadian Journal of Fisheries and Aquatic Sciences, 48, 734–750.CrossRefGoogle Scholar
Clark, W. G. (2002). F35% revisited ten years later. North American Journal of Fisheries Management, 22, 251–257.2.0.CO;2>CrossRefGoogle Scholar
Cochrane, K. L. (2000). Reconciling sustainability, economic efficiency and equity in fisheries: the one that got away?Fish and Fisheries, 1, 3–21.CrossRefGoogle Scholar
Codeço, C. T. (2000). Endemic and epidemic dynamics of cholera: the role of the aquatic reservoir. BMC Infectious Diseases, 1, 1–14.CrossRefGoogle Scholar
Cohen, D. (1966). Optimizing reproduction in a randomly varying environment. Journal of Theoretical Biology, 12, 119–129.CrossRefGoogle Scholar
Cohen, J. E. (1995). Unexpected dominance of high frequencies in chaotic nonlinear population models. Nature, 378, 610–612.CrossRefGoogle ScholarPubMed
Cole-King, A. (1993). Marine conservation: a new policy area. Marine Policy, May 1993, 171–185.CrossRefGoogle Scholar
Comins, H. N. and Hassell, M. P. (1979). The dynamics of optimally foraging predators and parasitoids. Journal of Animal Ecology, 48, 335–351.CrossRefGoogle Scholar
Comins, H. N., Hassell, M. P. and May, R. M. (1992). The spatial dynamics of host-parasitoid systems. Journal of Animal Ecology, 61, 735–748.CrossRefGoogle Scholar
Congdon, P. (2001). Bayesian Statistical Modelling. New York: John Wiley and Sons.Google Scholar
Conover, D. O. and Munch, S. B. (2002). Sustaining fisheries yields over evolutionary time scales. Science, 297, 94–96.CrossRefGoogle ScholarPubMed
Cook, R. M. and Hubbard, S. F. (1977). Adaptive searching strategies in insect parasites. Journal of Animal Ecology, 46, 115–125.CrossRefGoogle Scholar
Cooper, A. B., Hilborn, R. and Unsworth, J. W. (2003). An approach for population assessment in the absence of abundance indices. Ecological Applications, 13, 814–828.CrossRefGoogle Scholar
Cootner, P. H. (editor) (1964). The Random Character of Stock Market Prices. Cambridge, MA: MIT Press.Google Scholar
Corkett, C. J. (2002). Fish stock assessment as a non-falsifiable science: replacing an inductive and instrumental view with a critical rational one. Fisheries Research, 56, 117–123.CrossRefGoogle Scholar
Costantino, R. F. and Desharnais, R. A. (1991). Population Dynamics and the Tribolium Model: Genetics and Demography. New York: Springer Verlag.CrossRefGoogle Scholar
Cote, I. M., Mosqueira, I. and Reynolds, J. D. (2001). Effects of marine reserve characteristics on the protection of fish populations: a meta-analysis. Journal of Fish Biology, 59 (Supplement A), 178–179.CrossRefGoogle Scholar
Cottingham, K. L., Chiavelli, D. A. and Taylor, R. K. (2003). Environmental microbe and human pathogen: the ecology and microbiology of Vibrio cholerae. Frontiers in Ecology and the Environment, 1, 80–86.CrossRefGoogle Scholar
Council, N. R. (editor) (2001). Marine Protected Areas. Tools for Sustaining Ocean Ecosystems. Washington, DC: National Academy Press.Google Scholar
Courant, R. and Hilbert, D. (1962). Methods of Mathematical Physics. Volume II. Partial Differential Equations. New York: Wiley Interscience.Google Scholar
Crank, J. (1975). The Mathematics of Diffusion, 2nd edn. Oxford: Oxford University Press.Google Scholar
Crawley, M. J. (2002). Statistical Computing. An Introduction to Data Analysis Using S-Plus. New York: Wiley.Google Scholar
Crow, J. F. and Kimura, M. (1970). An Introduction to Population Genetics Theory. Minneapolis, MN: Burgess Publishing Company.Google Scholar
Crowder, L. B. and Murawski, S. A. (1998). Fisheries bycatch: Implications for management. Fisheries, 23, 8–17.2.0.CO;2>CrossRefGoogle Scholar
Daskalov, G. M. (2002). Overfishing drives a trophic cascade in the Black Sea. Marine Ecology – Progress Series, 225, 53–63.CrossRefGoogle Scholar
Davies, R. (1992). Reading and Writing. Salt Lake City: University of Utah Press.Google Scholar
Davis, H. T. (1962). Introduction to Nonlinear Differential and Integral Equations. New York: Dover.Google Scholar
Davis, P. J. and Gregerman, R. I. (1995). Occasional Notes: Parse Analysis II: A revised model that accounts for phi. The New England Journal of Medicine, 332, 965–966.CrossRefGoogle Scholar
Day, T. (2001). Parasite transmission models and the evolution of virulence. Evolution, 55, 2389–2400.CrossRefGoogle ScholarPubMed
Day, T. (2002a). The evolution of virulence in vector-borne and directly transmitted parasites. Theoretical Population Biology, 62, 199–213.CrossRefGoogle Scholar
Day, T. (2002b). On the evolution of virulence and the relationship between various measures of mortality. Proceedings of the Royal Society of London, B269, 1317–1323.CrossRefGoogle Scholar
Day, T. (2002c). Virulence evolution via host exploitation and toxin production in spore-producing pathogens. Ecology Letters, 5, 471–476.CrossRefGoogle Scholar
Day, T. (2003). Virulence evolution and the timing of disease life-history events. Trends in Ecology and Evolution, 18, 113–118.CrossRefGoogle Scholar
Day, T. and Proulx, S. (2004). A general theory for the evolutionary dynamics of virulence. American Naturalist, 164, E40–E63.CrossRefGoogle Scholar
Bach, P. (1943). The importance of host-feeding by adult parasites in the reduction of host populations. Journal of Economic Entomology, 36, 647–658.CrossRefGoogle Scholar
Bruijn, N. G. (1981). Asymptotic Methods in Analysis. New York: Dover Publications.Google Scholar
Grey, A. D. N. J. (2003a). An engineer's approach to the development of real anti-aging medicine. Science SAGE KE 2003 (8 January 2003).Google Scholar
Grey, A. D. N. J. (2003b). Overzealous maximum-likelihood fitting falsely convicts the slope heterogeneity hypothesis. Experimental Gerontology, 38, 921–923.CrossRefGoogle Scholar
Koeijer, A., Diekmann, O. and Reijnders, P. (1998). Modelling the spread of phocine distemper virus among harbour seals. Bulletin of Mathematical Biology, 60, 585–596.CrossRefGoogle ScholarPubMed
Koeijer, A., Heesterbeek, H., Schreuder, B., et al. (2004). Quantifying BSE control by calculating the basic reproduction ratio R0 for the infection among cattle. Journal of Mathematical Biology, 48, 1–22.CrossRefGoogle ScholarPubMed
Roode, J. C. and Read, A. F. (2003). Evolution and ecology, after the malaria genomes. Trends in Ecology and Evolution, 18, 60–61.CrossRefGoogle Scholar
Valpine, P. and Hastings, A. (2002). Fitting population models incorporating process noise and observation error. Ecological Monographs, 72, 57–76.CrossRefGoogle Scholar
DeAngelis, D. L. and Gross, L. J. (1992). Individual-Based Models and Approaches in Ecology. New York: Chapman and Hall.CrossRefGoogle Scholar
DeGroot, M. H. (1970). Optimal Statistical Decisions. New York: McGraw-Hill.Google Scholar
Delbrück, M. (1940). Statistical fluctuations in autocatalytic reactions. Journal of Chemical Physics, 8, 120–124.CrossRefGoogle Scholar
Demetrius, L. (2001). Mortality plateaus and directionality theory. Proceedings of the Royal Society of London, B268, 2029–2037.CrossRefGoogle ScholarPubMed
Denney, N. H., Jennings, S. and Reynolds, J. D. (2002). Life-history correlates of maximum population growth rates in marine fishes. Proceedings of the Royal Society of London, B269, 2229–2237.CrossRefGoogle ScholarPubMed
Dennis, B., Munholland, P. L. and Scott, J. M. (1991). Estimation of growth and extinction parameters for endangered species. Ecological Monographs, 61, 115–143.CrossRefGoogle Scholar
Dennis, B. and Otten, M. R. M. (2000). The joint effects of density dependence and rainfall on abundance of San Joaquin kit fox. Journal of Wildlife Management, 64, 388–400.CrossRefGoogle Scholar
Dick, E. J. (2004). Beyond ‘lognormal versus gamma’: discrimination among error distributions for generalized linear models. Fisheries Research, 70, 351–366.CrossRefGoogle Scholar
Dieckmann, U., Metz, J. A. J., Sabelis, M. W. and Sigmund, K. (2002). Adaptive dynamics of infectious diseases. In Pursuit of Virulence Management. Cambridge: Cambridge University Press.Google Scholar
Diserud, O. H. and Engen, S. (2000). A general and dynamic species abundance model, embracing the lognormal and gamma models. The American Naturalist, 155, 497–511.CrossRefGoogle ScholarPubMed
Dixit, A. K. and Pindyck, R. S. (1994). Investment Under Uncertainty. Princeton, NJ: Princeton University Press.Google Scholar
Dobson, A. and Foutopoulos, J. (2001). Emerging infectious pathogens of wildlife. Philosophical Transactions of the Royal Society London, B356, 1001–1012.CrossRefGoogle ScholarPubMed
Dobson, A. P. and Hudson, P. J. (1992). Regulation and stability of a free-living host–parasite system: Trichostrongylus tenius in red grouse. II. Population models. Journal of Animal Ecology, 61, 487–498.CrossRefGoogle Scholar
Dorn, M. W. (2002). Advice on west coast rockfish harvest rates from Bayesian meta-analysis of stock–recruit relationships. North American Journal of Fisheries Management, 22, 280–300.2.0.CO;2>CrossRefGoogle Scholar
Dovers, S. R. and Handmer, J. W. (1995). Ignorance, the precautionary principle, and sustainability. Ambio, 24, 92–97.Google Scholar
Dovers, S. R., Norton, T. W. and Handmer, J. W. (1996). Uncertainty, ecology, sustainability and policy. Biodiversity and Conservation, 5, 1143–1167.CrossRefGoogle Scholar
Driessen, G. and Hemerik, L. (1992). The time and egg budget of Leptopilina clavipes, a parasitoid of Drosophila. Ecological Entomology, 17, 17–27.CrossRefGoogle Scholar
Dubins, L. E. and Savage, L. J. (1976). Inequalities for Stochastic Processes: How to Gamble if You Must. New York: Dover Publications.Google Scholar
Dwyer, G. and Elkinton, J. S. (1993). Using simple models to predict virus epizootics in gypsy moth populations. Journal of Animal Ecology, 62, 1–11.CrossRefGoogle Scholar
Dyson, F. (1999). The Sun, the Genome, and the Internet. New York: Oxford University Press.Google Scholar
Earn, D. J. D., Dushoff, J. and Levin, S. A. (2002). Ecology and evolution of the flu. Trends in Ecology and Evolution, 17, 334–340.CrossRefGoogle Scholar
Easterling, M. R. and Ellner, S. P. (2000). Dormancy strategies in a random environment: comparing structured and unstructured models. Evolutionary Ecology Research, 2, 387–407.Google Scholar
Ebert, D. and Bull, J. J. (2003). Challenging the trade-off model for the evolution of virulence: is virulence management feasible?Trends in Microbiology, 11, 15–20.CrossRefGoogle ScholarPubMed
Edelstein-Keshet, L. (1988). Mathematical Models in Biology. New York: Random House.Google Scholar
Edgar, G. and Barrett, N. S. (1999). Effects of the declaration of marine reserves on Tasmanian reef fishes, invertebrates and plants. Journal of Experimental Marine Biology and Ecology, 242, 107–144.CrossRefGoogle Scholar
Edwards, A. W. F. (1992). Likelihood, expanded edition. Baltimore: Johns Hopkins University Press.Google Scholar
Edwards, R. L. (1954). The effect of diet on egg maturation and resorption in Mormoniella vitripennis (Hymenoptera, Pteromalidae). Quarterly Journal of Microscopical Science, 95, 459–469.Google Scholar
Efron, B. (2005). Bayesians, frequentists, and scientists. Journal of the American Statistical Association, 100, 1–5.CrossRefGoogle Scholar
Eigen, M. (1996). Prionics or the kinetic basis of prion diseases. Biophysical Chemistry, 63, A1–A18.CrossRefGoogle ScholarPubMed
Einstein, A. (1956). Investigations on the Theory of Brownian Movement. New York: Dover Publications.Google Scholar
Einum, S. and Fleming, I. A. (2000). Highly fecund mothers sacrifice offspring survival to maximize fitness. Nature, 405, 565–567.CrossRefGoogle ScholarPubMed
Ellison, A. M. (1996). An introduction to Bayesian inference for ecological research and environmental decision-making. Ecological Applications, 6, 1036–1046.CrossRefGoogle Scholar
Ellison, A. M. (2004). Bayesian inference in ecology. Ecology Letters, 7, 509–520.CrossRefGoogle Scholar
Emlen, J. M. (1966). The role of time and energy in food preference. American Naturalist, 100, 611–617.CrossRefGoogle Scholar
Enderby, J. E. (1998). Sir Frederick Charles Frank (1911–98). Nature, 393, 314.CrossRefGoogle Scholar
Engen, S., Lande, R. and Saether, B.-E. (2002). The spatial scale of population fluctuations and quasi-extinction risk. American Naturalist, 160, 439–451.Google ScholarPubMed
Engen, S. and Saether, B.-E. (2000). Predicting the time of quasi-extinction for populations far below their carrying capacity. Journal of Theoretical Biology, 205, 649–658.CrossRefGoogle ScholarPubMed
Engen, S., Saether, B.-E. and Møller, A. P. (2001). Stochastic population dynamics and time to extinction of a declining population of barn swallows. Journal of Animal Ecology, 70, 789–797.CrossRefGoogle Scholar
Enquist, B. J., West, G. B., Charnov, E. L. and Brown, J. H. (1999). Allometric scaling of production and life-history variation in vascular plants. Nature, 401, 907–911.CrossRefGoogle Scholar
Enserink, M. (2002). How devastating would a smallpox attack really be?Science, 296, 1592–1595.CrossRefGoogle Scholar
Enserink, M. (2003). New look at old data irks smallpox-eradication experts. Science, 299, 181.CrossRefGoogle Scholar
Esch, G. W., Shostak, A. W., Marcogliese, D. J. and Goater, T. M. (1990). Parasite Communities: Patterns and Processes. New York: Chapman and Hall.Google Scholar
Essington, T. E., Kitchell, J. F. and Walters, C. J. (2001). The von Bertalanffy growth function, bioenergetics, and the consumption rates of fish. Canadian Journal of Fisheries and Aquatic Sciences, 58, 2129–2138.CrossRefGoogle Scholar
Estes, J. A., Tinker, M. T., Williams, T. M. and Doak, D. F. (1998). Killer whale predation on sea otters linking oceanic and nearshore ecosystems. Science, 282, 473–476.CrossRefGoogle ScholarPubMed
Esteva, L. and Vargas, C. (2003). Coexistence of different serotypes of dengue virus. Journal of Mathematical Biology, 46, 31–47.CrossRefGoogle ScholarPubMed
Ewald, P. W. (1994). Evolution of Infectious Disease. New York: Oxford University Press.Google Scholar
Fanshawe, S., Vanblaricom, G. R. and Shelly, A. A. (2003). Restored top carnivores as detriments to the performance of marine protected areas intended for fishery sustainability: a case study with red abalones and sea otters. Conservation Biology, 17, 273–283.CrossRefGoogle Scholar
Farrow, S. and Sumaila, U. R. (2002). Economics of marine protected areas. Fish and Fisheries, 3, 356–359.Google Scholar
Faruque, S., Chowdhury, N., Kamruzzaman, M.et al. (2003). Reemergence of Epidemic Vibrio cholerae O139, Bangladesh. Emerging Infectious Diseases, 9, 1116–1122.CrossRefGoogle ScholarPubMed
Feller, W. (1957). An Introduction to Probability Theory and Its Applications, Volume I. New York: John Wiley and Sons.Google Scholar
Feller, W. (1971). An Introduction to Probability Theory and Its Applications, Volume II. New York: John Wiley and Sons.Google Scholar
Ferguson, H. M. and Read, A. F. (2002). Genetic and environmental determinants of malaria parasite virulence in mosquitoes. Proceedings of the Royal Society of London, B269, 1217–1224.CrossRefGoogle ScholarPubMed
Ferguson, N. M., Ghani, A. C., Donnelly, C. A., Hagenaars, T. J. and Anderson, R. M. (2002). Estimating the human health risk from possible BSE infection of the British sheep flock. Nature, 415, 420–424.CrossRefGoogle ScholarPubMed
Feynman, R. P. (1948). Space-time approach to non-relativistic quantum mechanics. Reviews of Modern Physics, 20, 367–387.CrossRefGoogle Scholar
Feynman, R. P. (1985). Surely You're Joking Mr. Feynman. New York: W. W. Norton.Google Scholar
Fieberg, J. and Ellner, S. P. (2000). When is it meaningful to estimate an extinction probability?Ecology, 81, 2040–2047.CrossRefGoogle Scholar
Fiksen, Ø. and Slotte, A. (2002). Stock-environment recruitment models for Norwegian spring spawning herring (Clupea harengus). Canadian Journal of Fisheries and Aquatic Sciences, 59, 211–217.CrossRefGoogle Scholar
Fisher, A. C., Hanemann, W. M. and Keeler, A. G. (1991). Integrating fishery and water resource management: a biological model of a California salmon fishery. Journal of Environmental Economics and Management, 20, 234–261.CrossRefGoogle Scholar
Fisher, R. A. (1930) (1958 reprint). The Genetical Theory of Natural Selection, 2nd revised edn. New York: Dover.CrossRefGoogle Scholar
Fishman, M. A. and Perelson, A. (1999). Th1/Th2 differentiation and cross-regulation. Bulletin of Mathematical Biology, 61, 403–436.CrossRefGoogle ScholarPubMed
Flanders, S. E. (1950). Regulation of ovulation and egg disposal in the parasitic hymenoptera. The Canadian Entomologist, 82, 134–140.CrossRefGoogle Scholar
Flather, C. H. (1996). Fitting species-accumulation functions and assessing regional land use impacts on avian diversity. Journal of Biogeography, 23, 155–168.CrossRefGoogle Scholar
Fletcher, J. P., Hughes, J. P. and Harvey, I. F. (1994). Life expectancy and egg load affect oviposition decision of a solitary parasitoid. Proceedings of the Royal Society of London, B258, 163–167.CrossRefGoogle Scholar
Foley, P. (1994). Predicting extinction times from environmental stochasticity and carrying capacity. Conservation Biology, 8, 124–137.CrossRefGoogle Scholar
Francis, R. I. C. C. and Shotton, R. (1997). “Risk” in fisheries management: a review. Canadian Journal of Fisheries and Aquatic Sciences, 54, 1699–1715.Google Scholar
Frank, F. C. (1953). On spontaneous asymmetric synthesis. Biochimica et Biophysica Acta, 11, 459–463.CrossRefGoogle ScholarPubMed
Frank, S. A. (1996). Models of parasite virulence. The Quarterly Review of Biology, 71, 37–78.CrossRefGoogle ScholarPubMed
Frank, S. A. (1998). Social Evolution. Princeton, NJ: Princeton University Press.Google ScholarPubMed
Frank, S. A. (2002). Immunology and Evolution of Infectious Diseases. Princeton, NJ: Princeton University Press.Google Scholar
Freedman, D., Pisani, R. and Purves, R. (1998). Statistics, 3rd edn. New York: W. W. Norton.Google Scholar
Friedlin, M. I. and Wentzell, A. D. (1984). Random Perturbations of Dynamical Systems. New York: Springer Verlag.CrossRefGoogle Scholar
Frisk, M. G., Miller, T. J. and Fogarty, M. J. (2001). Estimation and analysis of biological parameters in elasmobranch fishes: a comparative life history study. Canadian Journal of Fisheries and Aquatic Sciences, 58, 969–981.CrossRefGoogle Scholar
Fuerth, R. (1956). Notes. In Investigations on the Theory of the Brownian Movement by Albert Einstein, Ph.D. New York: Dover Publications, pp. 86–119.Google Scholar
Fulford, G. R., Roberts, M. G. and Heesterbeek, J. A. P. (2002). The metapopulation dynamics of an infectious disease: tuberculosis in possums. Theoretical Population Biology, 61, 15–29.CrossRefGoogle ScholarPubMed
Galvani, A. P. (2003). Epidemiology meets evolutionary ecology. Trends in Ecology and Evolution, 18, 132–139.CrossRefGoogle Scholar
Gandon, S., Mackinnon, M. J., Nee, S. and Read, A. F. (2001). Imperfect vaccines and the evolution of pathogen virulence. Nature, 414, 751–755.CrossRefGoogle ScholarPubMed
Gani, R. and Leach, S. (2001). Transmission potential of smallpox in contemporary populations. Nature, 414, 748–751.CrossRefGoogle ScholarPubMed
Ganusov, V. V., Bergstrom, C. T. and Antia, R. (2002). Within-host population dynamics and the evolution of microparasites in a heterogeneous host population. Evolution, 56, 213–223.CrossRefGoogle Scholar
Gardiner, C. W. (1983). Handbook of Stochastic Methods for Physics, Chemistry and the Natural Sciences. Berlin: Springer Verlag.CrossRefGoogle Scholar
Gardner, S. N. (2000). Scheduling chemotherapy: Catch 22 between cell kill and resistance evolution. Journal of Theoretical Medicine, 2, 21–232.CrossRefGoogle Scholar
Gardner, S. N. (2001). Modeling multi-drug chemotherapy: tailoring treatment to individuals. Journal of Theoretical Biology, 214, 181–207.CrossRefGoogle Scholar
Gardner, S. N. and Agrawal, A. A. (2002). Induced plant defense and the evolution of counter-defenses in herbivores. Evolutionary Ecology Research, 4, 1131–1151.Google Scholar
Gardner, S. N. and Thomas, M. B. (2002). Costs and benefits of fighting infection in locusts. Evolutionary Ecology Research, 4, 109–131.Google Scholar
Gavrilets, S. (2003). Models of speciation: what have we learned in 40 years?Evolution, 57, 2197–2215.CrossRefGoogle ScholarPubMed
Gavrilov, L. A. and Gavrilova, N. S. (1991). The Biology of Life Span: A Quantitative Approach. London: Harwood Academic Publishers.Google Scholar
Gelman, A., Carlin, J. B., Stern, H. S. and Rubin, D. B. (1995). Bayesian Data Analysis. London: Chapman and Hall.Google Scholar
George, E. (2004). Write Away. New York: Harper Collins.Google Scholar
Gillespie, D. T. (1992). Markov Processes. An Introduction for Physical Scientists. Boston: Academic Press.Google Scholar
Gillespie, D. T. (1996). Exact numerical simulation of the Ornstein–Uhlenbeck process and its integral. Physical Review E, 54, 2084–2091.CrossRefGoogle ScholarPubMed
Gillespie, J. H. (1991). The Causes of Molecular Evolution. New York: Oxford University Press.Google Scholar
Gillespie, J. H. (1998). Population Genetics. A Concise Guide. Baltimore: Johns Hopkins University Press.Google Scholar
Gillig, D., Griffin, W. L. and Ozuna, T. J. (2001). A bioeconomic assessment of Gulf of Mexico red snapper management policies. Transactions of American Fisheries Society, 130, 117–129.2.0.CO;2>CrossRefGoogle Scholar
Gillis, D. M. (1999). Behavioral inferences from regulatory observer data: catch rate variation in the Scotian Shelf silver hake (Merluccius bilinearis) fishery. Canadian Journal of Fisheries and Aquatic Sciences, 56, 288–296.CrossRefGoogle Scholar
Gillis, D. M. (2003). Ideal free distributions in fleet dynamics: a behavioral perspective on vessel movement in fisheries analysis. Canadian Journal of Zoology, 81, 177–187.CrossRefGoogle Scholar
Gillis, D. M. and Peterman, R. M. (1998). Implications of interference among fishing vessels and the ideal free distribution to the interpretation of CPUE. Canadian Journal of Fisheries and Aquatic Sciences, 55, 37–46.CrossRefGoogle Scholar
Gillis, D. M., Pikitch, E. K. and Peterman, R. M. (1995). Dynamic discarding decisions: foraging theory for high-grading in a trawl fishery. Behavioral Ecology, 6, 146–154.CrossRefGoogle Scholar
Giron, D., Rivero, A., Mandon, N., Darrouzet, E. and Casas, J. (2002). The physiology of host feeding in parasitic wasps: implications for survival. Functional Ecology, 16, 750–757.CrossRefGoogle Scholar
Giske, J., Mangel, M., Jakobsen, P., et al. (2002). Explicit trade-off rules in proximate adaptive agents. Evolutionary Ecology Reseach, 5, 835–865.Google Scholar
Gislason, H. (1994). Ecosystem effects of fishing activities in the North Sea. Marine Pollution Bulletin, 29, 520–527.CrossRefGoogle Scholar
Gleick, J. (1988). Chaos. London: William Heinneman Ltd.Google Scholar
Gleick, J. (1992). Genius. The Life and Science of Richard Feynman. New York: Pantheon.Google Scholar
Godfray, H. C. J. (1994). Parasitoids. Behavioral and Evolutionary Ecology. Princeton, NJ: Princeton University Press.Google Scholar
Godfray, H. C. J. and Hassell, M. P. (1989). Discrete and continuous insect populations in tropical environments. Journal of Animal Ecology, 58, 153–174.CrossRefGoogle Scholar
Goel, N. S. and Richter-Dyn, N. (1974). Stochastic Models in Biology. New York: Academic Press.Google Scholar
Gompertz, B. (1825). On the nature of the function expressive of the law of human mortality, and on a new mode of determining the value of life contingencies. Philosophical Transactions of the Royal Society, 115, 513–583.CrossRefGoogle Scholar
Gordon, D. M., Nisbet, R. M., Roos, A., Gurney, W. S. C. and Stewart, R. K. (1991). Discrete generations in host–parasitoid models with contrasting life cycles. Journal of Animal Ecology, 60, 295–308.CrossRefGoogle Scholar
Gordon, H. S. (1954). The economic theory of a common property resource: the fishery. Journal of Political Economy, 62, 124–142.CrossRefGoogle Scholar
Gotelli, N. J. (2001). A Primer of Ecology, 3rd edn. Sunderland, MA: Sinauer Associates.Google Scholar
Gould, S. J. (2002). The Structure of Evolutionary Theory. Cambridge, MA: The Belknap Press of Harvard University Press.Google Scholar
Gould, S. J. and Eldredge, N. (1977). Punctuated equilibria: the tempo and mode of evolution reconsidered. Paleobiology, 3, 115–151.CrossRefGoogle Scholar
Gould, S. J. and Eldredge, N. (1993). Punctuated equilibrium comes of age. Nature, 366, 223–227.CrossRefGoogle ScholarPubMed
Grafen, A. and Hails, R. S. (2002). Modern Statistics for the Life Sciences. Learn How to Analyse your Experiments. Oxford: Oxford University Press.Google Scholar
Graham, A. L. (2001). Use of an optimality model to solve the immunological puzzle of concomitant infection. Parasitology, 122, S61–S64.CrossRefGoogle ScholarPubMed
Graham, A. L. (2002). When T-helper cells don't help: immunopathology during concomitant infection. Quarterly Review of Biology, 77, 409–434.CrossRefGoogle ScholarPubMed
Gray, R. D. (1987). Faith and foraging: a critique of the “paradigm argument from design”. In Kamil, A. C., Krebs, J. R. and Pulliam, H. R. (editors) Foraging Behavior. New York: Plenum, pp. 69–140.CrossRefGoogle Scholar
Greenberg, R. (2001). Lecture 38: Nineteenth-Century Italian Opera – Giuseppe Verdi. In How to Listen to and Understand Great Music. The Teaching Company.Google Scholar
Greene, C. (2000). Habitat selection reduces extinction of populations subject to Allee effects. Theoretical Population Biology, 64, 1–10.CrossRefGoogle Scholar
Greenwood, M. and Yule, G. U. (1920). An inquiry into the nature of frequency distributions representative of multiple happenings with particular references to the occurrence of multiple attacks of disease or of repeated accidents. Journal of the Royal Statistical Society, 83, 255–279.CrossRefGoogle Scholar
Greer, R. (1995). Ferox Trout and Arctic Charr. A Predator, its Pursuit and its Prey. Shrewsbury, UK: Swan Hill Press.Google Scholar
Grenfell, B. T. (1988). Gastrointestinal nematode parasites and the stability and productivity of intensive ruminant grazing systems. Philosophical Transactions of the Royal Society of London, B321, 541–563.CrossRefGoogle ScholarPubMed
Grenfell, B. T. (1992). Parasitism and the dynamics of ungulate grazing systems. American Naturalist, 139, 907–929.CrossRefGoogle Scholar
Grindrod, P. (1996). The Theory and Applications of Reaction–Diffusion Equations. Patterns and Waves. Oxford: Oxford University Press.Google Scholar
Groot, C. and Margolis, L. (editors) (1991). Pacific Salmon Life Histories. Vancouver, BC: University of British Columbia Press.Google Scholar
Groot, C., Margolis, L. and Clarke, W. C. (editors) (1995). The Physiological Ecology of Pacific Salmon. Vancouver, BC: University of British Columbia Press.Google Scholar
Guedj, D. (2000). The Parrot's Theorem. New York: St. Martin's Griffin.Google Scholar
Guenette, S. and Pitcher, T. J. (1999). An age-structured model showing the benefits of marine reserves in controlling overexploitation. Fisheries Research, 39, 295–303.CrossRefGoogle Scholar
Guenette, S., Lauck, T. and Clark, C. (1998). Marine reserves: from Beverton and Holt to the present. Reviews in Fish Biology and Fisheries, 8, 251–272.CrossRefGoogle Scholar
Gulland, J. A. (editor) (1977). Fish Population Dynamics. Chichester: John Wiley and Sons.Google Scholar
Gulland, J. A. (1988). Fish Population Dynamics. Implications for Management, 2nd edn. Chichester: John Wiley & Sons.Google Scholar
Gunderson, D. R. (1997). Trade-off between reproductive effort and adult survival in oviparous and viviparous fishes. Canadian Journal of Fisheries and Aquatic Sciences, 54, 990–998.CrossRefGoogle Scholar
Gurney, W. S. C., Nisbet, R. M. and Lawton, J. H. (1983). The systematic formulation of tractable single-species population models incorporating age structure. Journal of Animal Ecology, 52, 479–495.CrossRefGoogle Scholar
Haberman, R. (1998). Elementary Applied Partial Differential Equations with Fourier Series and Boundary Value Problems. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Haccou, P., Vlas, S. J., Alhen, J. J. M. and Visser, M. E. (1991). Information processing by foragers: effects of intra-patch experience on the leaving tendency of Leptopilina heterotoma. Journal of Animal Ecology, 60, 93–106.CrossRefGoogle Scholar
Hadamard, J. (1954). The Psychology of Invention in the Mathematical Field. New York: Dover.Google Scholar
Hakoyama, H. and Iwasa, Y. (2000). Extinction risk of a density-dependent population estimated from a time series of population size. Journal of Theoretical Biology, 204, 337–359.CrossRefGoogle ScholarPubMed
Hakoyama, H., Iwasa, Y. and Nakanishi, J. (2000). Comparing risk factors for population extinction. Journal of Theoretical Biology, 204, 327–336.CrossRefGoogle ScholarPubMed
Halley, J. M. and Iwasa, Y. (1998). Extinction rate of a population under both demographic and environmental stochasticity. Theoretical Population Biology, 53, 1–15.CrossRefGoogle ScholarPubMed
Halpern, B. S. and Warner, R. R. (2002). Marine reserves have rapid and lasting effects. Ecology Letters, 5, 361–366.CrossRefGoogle Scholar
Ham, K. D. and Pearsons, T. N. (2000). Can reduced salmonid population abundance be detected in time to limit management impacts?Canadian Journal of Fisheries and Aquatic Sciences, 57, 17–24.CrossRefGoogle Scholar
Hamilton, W. D. (1966). The moulding of senescence by natural selection. Journal of Theoretical Biology, 12, 12–45.CrossRefGoogle ScholarPubMed
Hamilton, W. D. (1967). Extraordinary sex ratios. Science, 156, 477–488.CrossRefGoogle ScholarPubMed
Hamilton, W. D. (1995). Narrow Roads of Gene Land. Volume 1. Evolution of Social Behaviour. New York: W. H. Freeman and Company.Google Scholar
Hammersley, J. M. (1974). Statistical tools. The Statistician, 23, 89–106.CrossRefGoogle Scholar
Hammond, T. R. and O'Brien, C. M. (2001). An application of the Bayesian approach to stock assessment model uncertainty. ICES Journal of Marine Science, 58, 648–656.CrossRefGoogle Scholar
Hanski, I. (1989). Metapopulation dynamics: does it help to have more of the same?Trends in Ecology and Evolution, 4, 113–114.CrossRefGoogle ScholarPubMed
Hanski, I. (1999). Metapopulation Ecology. Oxford: Oxford University Press.Google Scholar
Hare, S. R. and Francis, R. C. (1995). Climate change and salmon production in the Northeast Pacific Ocean. Canadian Special Publications of Fisheries and Aquatic Sciences, 121, 357–372.Google Scholar
Harley, S. J. and Myers, R. A. (2001). Hierarchial Bayesian models of length-specific catchability of research trawl surveys. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1569–1584.CrossRefGoogle Scholar
Harley, S. J., Myers, R. A. and Dunn, A. (2001). Is catch-per-unit effort proportional to abundance?Canadian Journal of Fisheries and Aquatic Sciences, 58, 1760–1772.CrossRefGoogle Scholar
Harris, P. J. and Dean, J. M. (1998). Characterization of king mackerel and Spanish mackerel bycatches of South Carolina shrimp trawlers. North American Journal of Fisheries Management, 18, 439–453.2.0.CO;2>CrossRefGoogle Scholar
Harrison, L. G. (1993). Kinetic Theory of Living Pattern. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Hart, P. J. B. and Reynolds, J. D. (editors) (2002). Handbook of Fish Biology and Fisheries, Volume 2. Fisheries. Oxford: Blackwell.CrossRefGoogle Scholar
Harte, J. (1988). Consider a Spherical Cow. A Course in Environmental Problem Solving. Sausalito, CA: University Science Books.Google Scholar
Harte, J. (2001). Consider a Cylindrical Cow. More Adventures in Environmental Problem Solving. Sausalito, CA: University Science Books.Google Scholar
Hassell, M. P. (1978). The Dynamics of Arthropod Predator Prey Systems. Princeton, NJ: Princeton University Press.Google ScholarPubMed
Hassel, M. P. (2000a). The Spatial and Temporal Dynamics of Host–Parasitoid Interactions. Oxford: Oxford University Press.Google Scholar
Hassell, M. P. (2000b). Host–parasitoid population dynamics. Journal of Animal Ecology, 69, 543–566.CrossRefGoogle Scholar
Hassell, M. P. and May, R. M. (1988). Spatial heterogeneity and the dynamics of parasitoid–host systems. Annals Zoologia Fennici, 25, 55–61.Google Scholar
Hassell, M. P., Waage, J. K. and May, R. M. (1983). Variable parasitoid sex ratios and their effect on host-parasitoid dynamics. Journal of Animal Ecology, 52, 889–904.CrossRefGoogle Scholar
Hastings, A. and Botsford, L. W. (1999). Equivalence in yield from marine reserves and traditional fisheries management. Science, 284, 1537–1538.CrossRefGoogle ScholarPubMed
Hastings, I. M. (1997). A model for the origins and spread of drug-resistant malaria. Parasitology, 115, 133–141.CrossRefGoogle ScholarPubMed
Hastings, I. M. (2001). Modelling parasite drug resistance: lessons for management and control strategies. Tropical Medicine and International Health, 6, 883–890.CrossRefGoogle ScholarPubMed
Hastings, I. M. and D'Allesandro, U. (2000). Modelling a predictable disaster: the rise and spread of drug-resistant malaria. Parasitology Today, 16, 340–347.CrossRefGoogle Scholar
Hastings, I. M. and MacKinnon, M. J. (1998). The emergence of drug-resistant malaria. Parasitology, 117, 411–417.CrossRefGoogle ScholarPubMed
Hastings, I. M., Bray, P. G. and Ward, S. A. (2002). A requiem for chloroquine. Science, 297, 74–75.CrossRefGoogle Scholar
Haupt, R. L. and Haupt, S. E. (1998). Practical Genetic Algorithms. New York: John Wiley and Sons.Google Scholar
He, J. X. and Stewart, D. J. (2001). Age and size at first reproduction of fishes: predictive models based only on growth trajectories. Ecology, 82, 784–791.CrossRefGoogle Scholar
Healey, M. C. (1985). Influence of fishermen's preference on the success of commercial fishery management regimes. North American Journal of Fisheries Management, 5, 173–180.2.0.CO;2>CrossRefGoogle Scholar
Healey, M. C. (1990). Implications of climate change for fisheries management policy. Transactions of American Fisheries Society, 119, 366–373.2.3.CO;2>CrossRefGoogle Scholar
Healey, M. C. and Morris, J. F. T. (1992). The relationship between the dispersion of salmon fishing vessels and their catch. Fisheries Research, 15, 135–145.CrossRefGoogle Scholar
Heath, D. D., Heath, J. W., Bryden, C. A., Johnson, R. M. and Fox, C. W. (2003). Rapid evolution of egg size in captive salmon. Science, 299, 1738–1740.CrossRefGoogle ScholarPubMed
Hedrick, P. W., Lacy, R. C., Allendorf, F. W. and Soule, M. E. (1996). Directions in conservation biology. Conservation Biology, 10, 1312–1320.CrossRefGoogle Scholar
Heimpel, G. E. and Collier, T. R. (1996). The evolution of host-feeding behaviour in parasitoids. Biological Reviews, 71, 373–400.CrossRefGoogle Scholar
Heimpel, G. E. and Rosenheim, J. A. (1995). Dynamic host feeding by the parasitoid Aphytis melinus: the balance between current and future reproduction. Journal of Animal Ecology, 64, 153–167.CrossRefGoogle Scholar
Heimpel, G. E. and Rosenheim, J. A. (1998). Egg limitation in parasitoids: a review of the evidence and a case study. Biological Control, 11, 160–168.CrossRefGoogle Scholar
Heimpel, G. E., Mangel, M. and Rosenheim, J. A. (1998). Effects of time limitation and egg limitation on lifetime reproductive success of a parasitoid in the field. American Naturalist, 152, 273–289.CrossRefGoogle ScholarPubMed
Heimpel, G. E., Rosenheim, J. A. and Adams, J. A. (1994). Behavioral ecology of host feeding in Aphytis parasitoids. Norwegian Journal of Agricultural Sciences, Supplement, 16, 101–115.Google Scholar
Heimpel, G. E., Rosenheim, J. A. and Mangel, M. (1996). Egg limitation, host quality, and dynamic behavior by a parasitoid in the field. Ecology, 77, 2410–2420.CrossRefGoogle Scholar
Heimpel, G. E., Rosenheim, J. A. and Kattari, D. (1997). Adult feeding and lifetime reproductive success in the parasitoid Aphytis melinus. Entomologia Experimentalis et Applicata, 83, 305–315.CrossRefGoogle Scholar
Hellriegel, B. (2001). Immunoepidemiology – bridging the gap between immunology and epidemiology. Trends in Parasitology, 17, 102–106.CrossRefGoogle ScholarPubMed
Helser, T. E., Thunberg, E. M. and Mayo, R. K. (1996). An age-structured bioeconomic simulation of U.S. Silver hake Fisheries. North American Journal of Fisheries Management, 16, 783–794.2.3.CO;2>CrossRefGoogle Scholar
Hemerik, L., Driessen, G. and Haccou, P. (1993). Effects of intra-patch experiences on patch time, search time and searching efficiency of the parasitoid Leptopilina clavipes. Journal of Animal Ecology, 62, 33–44.CrossRefGoogle Scholar
Hemerik, L., Hoeven, N. and Alphen, J. J. M. (2002). Egg distributions and the information a solitary parasitoid has and uses for its oviposition decisions. Acta Biotheoretica, 50, 167–188.CrossRefGoogle ScholarPubMed
Henderson, D. A. (1999). The looming threat of bioterrorism. Science, 283, 1279–1282.CrossRefGoogle ScholarPubMed
Hendry, A. P. and Kinnison, M. T. (2001). An introduction to microevolution: rate, pattern, process. Genetica, 112–113, 1–8.CrossRefGoogle ScholarPubMed
Hennemuth, R. C., Palmer, J. E. and Brown, B. E. (1980). A statistical description of recruitment in eighteen selected fish stocks. Journal of Northwest Atlantic Fishery Science, 1, 101–111.CrossRefGoogle Scholar
Hernandez, M.-J. and Barradas, I. (2003). Variation in the outcome of population interactions: bifurcations and catastrophes. Journal of Mathematical Biology, 46, 571–594.CrossRefGoogle ScholarPubMed
Hershatter, G. (1997). Dangerous Pleasures. Berkeley, CA: University of California Press.Google Scholar
Hess, G. (1996). Disease in metapopulation models: implications for conservation. Ecology, 77, 1617–1632.CrossRefGoogle Scholar
Hethcote, H. W. (2000). The mathematics of infectious diseases. SIAM Review, 42, 599–653.CrossRefGoogle Scholar
Higham, D. J. (2001). An algorithmic introduction to numerical simulation of stochastic differential equations. SIAM Review, 43, 525–546.CrossRefGoogle Scholar
Highman, N. J. (1998). Handbook of Writing for the Mathematical Sciences, 2nd edn. Philadelphia, PA: SIAM (Society for Industrial and Applied Mathematics).Google Scholar
Hilborn, R. and Mangel, M. (1997). The Ecological Detective. Confronting Models with Data. Princeton, NJ: Princeton University Press.Google Scholar
Hilborn, R., Parama, A. and Maunder, M. (2002). Exploitation rate reference points for west coast rockfish: are they robust and are there better alternatives?North American Journal of Fisheries Management, 22, 365–375.2.0.CO;2>CrossRefGoogle Scholar
Hines, W. G. S. (1987). Evolutionarily stable strategies: a review of basic theory. Theoretical Population Biology, 31, 195–272.CrossRefGoogle ScholarPubMed
Hinrichsen, R. A. (2002). The accuracy of alternative stochastic growth rate estimates for salmon populations. Canadian Journal of Fisheries and Aquatic Sciences, 59, 1014–1023.CrossRefGoogle Scholar
Hochberg, M. E. and Ives, A. R. (editors) (2000). Parasitoid Population Biology. Princeton, NJ: Princeton University Press.Google Scholar
Hofbauer, J. and Sigmund, K. (1998). Evolutionary Games and Population Dynamics. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Holden, A. (editor) (2001). The New Penguin Opera Guide. London: Penguin Books.Google Scholar
Holland, D. S. and Sutinen, J. G. (1999). An empirical model of fleet dynamics in New England trawl fisheries. Canadian Journal of Fisheries and Aquatic Sciences, 56, 253–264.CrossRefGoogle Scholar
Holoman, D. K. (1992). Evenings with the Orchestra. New York: Norton.Google Scholar
Horwood, J. W., Nichols, J. H. and Milligan, S. (1998). Evaluation of closed areas for fish stock conservation. Journal of Applied Ecology, 35, 893–903.CrossRefGoogle Scholar
Hotelling, H. (1951). The impact of R. A. Fisher on statistics. Journal of the American Statistical Association, 46, 35–46.CrossRefGoogle Scholar
Houston, A. I. and McNamara, J. M. (1999). Models of Adaptive Behaviour. Cambridge: Cambridge University Press.Google Scholar
Hubbard, S. F. and Cook, R. M. (1978). Optimal foraging by parasitoid wasps. Journal of Animal Ecology, 47, 593–604.CrossRefGoogle Scholar
Hubbard, S. F., Marris, G. and Reynolds, D. M. (1987). Adaptive patterns in the avoidance of superparasitism by solitary parasitic wasps. Journal of Animal Ecology, 56, 387–401.CrossRefGoogle Scholar
Hubbell, S. (2001). The Unified Neutral Theory of Biodiversity and Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
Hudson, D. J. (1971). Interval estimation from the likelihood function. Journal of the Royal Statistical Society, Series B, 33, 256–262.Google Scholar
Hudson, P. J. (1986). The effect of a parasitic nematode on the breeding production of red grouse. Journal of Animal Ecology, 55, 85–92.CrossRefGoogle Scholar
Hudson, P. J. and Dobson, A. P. (1997). Transmission dynamics and host–parasite interactions of Trichostrongylus tenuis in red grouse (Lagopus lagopus scoticus). Journal of Parasitology, 83, 194–202.CrossRefGoogle Scholar
Hudson, P. J., Dobson, A. P. and Newborn, D. (1992a). Do parasites make prey vulnerable to predation – red grouse and parasites. Journal of Animal Ecology, 61, 681–692.CrossRefGoogle Scholar
Hudson, P. J., Newborn, D. and Dobson, A. P. (1992b). Regulation and stability of a free-living host–parasite system – Trichostrongylus tenuis in red grouse. 1. Monitoring and parasite reduction experiments. Journal of Animal Ecology, 61, 477–486.CrossRefGoogle Scholar
Hudson, P. J., Dobson, A. P. and Newborn, D. (1998). Prevention of population cycles by parasite removal. Science, 282, 2256–2258.CrossRefGoogle ScholarPubMed
Hudson, P. J., Dobson, A. P., Cattadori, I. M., et al. (2002). Trophic interactions and population growth rates: describing patterns and identifying mechanisms. Philosophical Transactions of the Royal Society of London, B357, 1259–1271.CrossRefGoogle ScholarPubMed
Huelsenbeck, J. P. and Ronquist, F. (2001). MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics, 17, 754–755.CrossRefGoogle ScholarPubMed
Huelsenbeck, J. P., Ronquist, F., Nielsen, R. and Bollback, J. P. (2001). Bayesian inference of phylogeny and its impact on evolutionary biology. Science, 294, 2310–2314.CrossRefGoogle ScholarPubMed
Hughes, B. D. (1995). Random Walks and Random Environments. Volume 1: Random Walks. Oxford: Clarendon Press.Google Scholar
Hur, K., Kim, J.-I., Choi, S.-I., et al. (2002). The pathogenic mechanisms of prion diseases. Mechanisms of Ageing and Development, 123, 1637–1647.CrossRefGoogle ScholarPubMed
Huse, G., Strand, E. and Giske, J. (1999). Implementing behaviour in individual-based models using neural networks and genetic algorithms. Evolutionary Ecology, 13, 469–483.CrossRefGoogle Scholar
Hutchings, J. A. (2000). Collapse and recovery of marine fishes. Nature, 406, 882–886.CrossRefGoogle ScholarPubMed
Hutchings, J. A. (2001). Influence of population decline, fishing, and spawner variability on the recovery of marine fishes. Journal of Fish Biology, 59 (Supplement A), 306–322.CrossRefGoogle Scholar
Hutchings, J. A. and Myers, R. A. (1994). What can be learned from the collapse of a renewable resource? Atlantic cod, Gadus morhua, of Newfoundland and Labrador. Canadian Journal of Fisheries and Aquatic Sciences, 51, 2126–2146.CrossRefGoogle Scholar
Hutchings, J. A., Walters, C. and Haedrich, R. L. (1997). Is scientific inquiry incompatible with government information control?Canadian Journal of Fisheries and Aquatic Sciences, 54, 1198.CrossRefGoogle Scholar
Iwasa, Y., Hakoyama, H., Nakamaru, M. and Nakanishi, J. (2000). Estimate of population extinction risk and its application to ecological risk management. Population Ecology, 42, 73–80.CrossRefGoogle Scholar
Jacobson, L. D. and MacCall, A. D. (1995). Stock–recruitment models for Pacific Sardine (Sardinops sagax). Canadian Journal of Fisheries and Aquatic Sciences, 52, 566–577.CrossRefGoogle Scholar
Jacobson, L. D., Oliveira, J. A. A., Barange, M., et al. (2001). Surplus production, variability, and climate change in the great sardine and anchovy fisheries. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1891–1903.CrossRefGoogle Scholar
Jacobson, L. D., Cadrin, S. X. and Weinberg, J. R. (2002). Tools for estimating surplus production and FMSY in any stock assessment model. North American Journal of Fisheries Management, 22, 326–338.2.0.CO;2>CrossRefGoogle Scholar
James, A., Pitchford, J. W. and Brindley, J. (2003). The relationship between plankton blooms, the hatching of fish larvae, and recruitment. Ecological Modelling, 160, 77–90.CrossRefGoogle Scholar
Janssen, A. (1989). Optimal host selection by Drosophila parasitoids in the field. Functional Ecology, 3, 469–479.CrossRefGoogle Scholar
Janssen, A., Driessen, G., Haan, M. and Roodbol, N. (1988). The impact of parasitoids on natural populations of temperature woodland Drosophila. Netherlands Journal of Zoology, 38, 61–73.CrossRefGoogle Scholar
Jaynes, E. T. (2003). Probability Theory. The Logic of Science. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Jenkins, R. (2001). Churchill. New York: Farrar, Straus and Grioux.Google Scholar
Jennings, S. (2000). Patterns and prediction of population recovery in marine reserves. Reviews in Fish Biology and Fisheries, 10, 209–231.CrossRefGoogle Scholar
Jennings, S. and Polunin, N. V. C. (1996). Fishing strategies, fishery development and socioeconomics in traditionally managed Fijian fishing grounds. Fisheries Management and Ecology, 3, 335–347.CrossRefGoogle Scholar
Jennings, S., Reynolds, J. D. and Mills, S. C. (1998). Life history correlates of responses to fisheries exploitation. Proceedings of the Royal Society of London, B265, 333–339.CrossRefGoogle Scholar
Jennings, S., Kaiser, M. and Reynolds, J. D. (2001). Marine Fisheries Ecology. Oxford: Blackwell Science.Google Scholar
Jervis, M. A. and Kidd, N. A. C. (1986). Host-feeding strategies in hymenopteran parasitoids. Biological Reviews, 61, 395–434.CrossRefGoogle Scholar
Jervis, M. A., Heimpel, G. E., Ferns, P. N., Harvey, J. A. and Kidd, N. A. C. (2001). Life-history strategies in parasitoid wasps: a comparative analysis of ‘ovigeny’. Journal of Animal Ecology, 70, 442–458.CrossRefGoogle Scholar
Jonzen, N., Cardinale, M., Gardmark, A., Arrhenius, F. and Lundberg, P. (2002). Risk of collapse in the eastern Baltic cod fishery. Marine Ecology – Progress Series, 240, 225–233.CrossRefGoogle Scholar
Kac, M. (1985). Enigmas of Chance. New York: Harper and Row.Google Scholar
Kailath, T. (1980). Linear Systems. Englewood Cliffs, NJ: Prentice Hall.Google Scholar
Kaplan, E. H. and Wein, L. M. (2003). Smallpox eradication in west and central Africa: surveillance-containment or herd immunity?Epidemiology, 14, 90–92.CrossRefGoogle ScholarPubMed
Kaplan, E. H., Craft, D. L. and Wein, L. M. (2002). Emergency response to a smallpox attack: the case for mass vaccination. Proceedings of the National Academy of Sciences, 99, 10 935–10 940.CrossRefGoogle ScholarPubMed
Karlin, S. and Taylor, H. M. (1981). A Second Course in Stochastic Processes. New York: Academic Press.Google Scholar
Keeling, M. J. and Grenfell, B. T. (1997). Disease extinction and community size: modeling the persistence of measles. Science, 275, 65–67.CrossRefGoogle ScholarPubMed
Keeling, M. J. and Grenfell, B. T. (2000). Individual-based perspectives on R0. Journal of Theoretical Biology, 203, 51–61.CrossRefGoogle Scholar
Kehler, D. G., Myers, R. A. and Field, C. A. (2002). Measurement error and bias in the maximum reproductive rate for the Ricker model. Canadian Journal of Fisheries and Aquatic Sciences, 59, 854–864.CrossRefGoogle Scholar
Keizer, J. E. (1987). Statistical Thermodynamics of Nonequilibrium Processes. New York: Springer Verlag.CrossRefGoogle Scholar
Keller, E. F. (2002). Making Sense of Life. Cambridge, MA: Harvard University Press.Google Scholar
Keller, J. B. (1974). Optimal velocity in a race. American Mathematical Monthly, 81, 474–480.CrossRefGoogle Scholar
Kendall, M. and Stuart, A. (1979). The Advanced Theory of Statistics. Volume 2. Inference and Relationship, 4th edn. London: Charles Griffin and Company Limited.Google Scholar
Kermack, W. O. and McKendrick, A. G. (1927). A contribution to the mathematical theory of epidemics. Proceedings of the Royal Society of London, A115, 700–721.CrossRefGoogle Scholar
Kermack, W. O. and McKendrick, A. G. (1932). Contributions to the mathematical theory of epidemics – II. The problem of endemicity. Proceedings of the Royal Society of London, A138, 55–83.CrossRefGoogle Scholar
Kermack, W. O. and McKendrick, A. G. (1933). Contributions to the mathematical study of epidemics. III. Further studies of the problem of endemicity. Proceedings of the Royal Society of London, A141, 94–122.CrossRefGoogle Scholar
Kimura, M. and Ohta, T. (1971). Theoretical Aspects of Population Genetics. Princeton, NJ: Princeton University Press.Google ScholarPubMed
Kinas, P. G. (1996). Bayesian fishery stock assessment and decision making using adaptive importance sampling. Canadian Journal of Fisheries and Aquatic Sciences, 53, 414–423.CrossRefGoogle Scholar
King, J. R. and McFarlane, G. A. (2003). Marine fish life history strategies: applications to fishery management. Fisheries Management and Ecology, 10, 249–264.CrossRefGoogle Scholar
King, M. (1995). Fisheries Biology, Assessment and Management. Ames, Iowa: Iowa State University Press.Google Scholar
King, S. (2000). On Writing. New York: Scribner.Google Scholar
Kingsland, S. E. (1985). Modeling Nature. Chicago: University of Chicago Press.Google Scholar
Kirchner, J. W. (2001). Fractal power spectra plotted upside-down. Comment on “Scaling of power spectrum of extinction events in the fossil record” by V. P. Dimri and M. R. Prakash. Earth and Planetary Science Letters, 192, 617–621.CrossRefGoogle Scholar
Kirchner, J. W. (2002). Evolutionary speed limits inferred from the fossil record. Nature, 415, 65–68.CrossRefGoogle ScholarPubMed
Kirchner, J. W. and Roy, B. A. (1999). The evolutionary advantages of dying young: epidemiological implications of longevity in metapopulations. American Naturalist, 154, 140–159.CrossRefGoogle ScholarPubMed
Kirchner, J. W. and Weil, A. (1998). No fractals in fossil extinction statistics. Nature, 395, 337–338.CrossRefGoogle Scholar
Kirchner, J. W. and Weil, A. (2000). Correlations in fossil extinction and origination rates through geological time. Proceedings of the Royal Society of London, B267, 1301–1309.CrossRefGoogle ScholarPubMed
Kirkwood, T. B. L. (1999). Evolution, molecular biology and mortality plateaus. In Bohr, V. A., Clark, B. F. C. and Stevensner, T. (editors) Molecular Biology of Aging, Alfred Benzon Symposium 44. Copenhagen: Munksgaard, pp. 383–390.Google Scholar
Klein, E. K., Lavigne, C., Foueillassar, X., Gouyon, P. and Larédo, C. (2003). Corn pollen dispersal: quasi-mechanistic models and field experiments. Ecological Monographs, 73, 131–150.CrossRefGoogle Scholar
Klein, G. (1952). Mean first-passage times of Brownian motion and related problems. Proceedings of the Royal Society of London, A211, 431–443.CrossRefGoogle Scholar
Klyashtorin, L. B. (1998). Long-term climate change and main commercial fish production in the Atlantic and Pacific. Fisheries Research, 37, 115–125.CrossRefGoogle Scholar
Knell, R. J., Begon, M. and Thompson, D. J. (1996). Transmission dynamics of Bacillus thuringiensis infecting Plodia interpunctella: a test of the mass action assumption with an insect pathogen. Proceedings of the Royal Society of London, B263, 75–81.CrossRefGoogle ScholarPubMed
Koella, J. C. and Boëte, C. (2003). A model for the coevolution of immunity and immune evasion in vector-borne diseases with implications for the epidemiology of malaria. The American Naturalist, 161, 698–707.CrossRefGoogle ScholarPubMed
Koella, J. C. and Restif, O. (2001). Coevolution of parasite virulence and host life history. Ecology Letters, 4, 207–214.CrossRefGoogle Scholar
Koeller, P. (2003). The lighter side of reference points. Fisheries Research, 62, 1–6.CrossRefGoogle Scholar
Kolata, G. B. (1977). Catastrophe theory: the emperor has no clothes. Science, 196, 287 + 350–351.CrossRefGoogle ScholarPubMed
Koopman, B. O. (1980). Search and Screening. General Principles with Historical Applications. Elmsford, NY: Pergamon Press.Google Scholar
Korman, J. and Higgins, P. S. (1997). Utility of escapement time series data for monitoring the response of salmon populations to habitat alteration. Canadian Journal of Fisheries and Aquatic Sciences, 54, 2058–2067.CrossRefGoogle Scholar
Kot, M. (2001). Elements of Mathematical Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Kot, M., Lewis, M. A. and Driessche, P. (1996). Dispersal data and the spread of invading organisms. Ecology, 77, 2027–2042.CrossRefGoogle Scholar
Kramers, H. A. (1940). Brownian motion in a field of force and the diffusion model of chemical reactions. Physica, 7, 284–312.CrossRefGoogle Scholar
Kubo, R., Matsuo, K. and Kitahara, K. (1973). Fluctuation and relaxation of macrovariables. Journal of Statistical Physics, 9, 51–96.CrossRefGoogle Scholar
Kuikka, S., Hilden, M., Gislason, H. F., et al. (1999). Modeling environmentally driven uncertainties in Baltic cod (Gadus morhua) management by Bayesian inference diagrams. Canadian Journal of Fisheries and Aquatic Sciences, 56, 629–641.CrossRefGoogle Scholar
Laakso, J., Kaitala, V. and Ranta, E. (2003). Non-linear biological responses to disturbance: consequences on population dynamics. Ecological Modelling, 162, 247–258.CrossRefGoogle Scholar
Lande, R. (1985). Expected time for random genetic drift of a population between stable phenotypic states. Proceedings of the National Academy of Sciences, 82, 7641–7645.CrossRefGoogle ScholarPubMed
Lande, R. (1987). Extinction thresholds in demographic models of territorial species. American Naturalist, 130, 624–635.CrossRefGoogle Scholar
Lande, R., Engen, S. and Saether, B.-E. (2003). Stochastic Population Dynamics in Ecology and Conservation. Oxford: Oxford University Press.CrossRefGoogle Scholar
Lander, A. D. (2004). A calculus of purpose. PLoS Biology, 2, 0712–0714.CrossRefGoogle ScholarPubMed
Laurenson, M. K., Norman, R. A., Gilbert, L., Reid, H. W. and Hudson, P. J. (2003). Identifying disease reservoirs in complex systems: mountain hares as reservoirs of ticks and louping-ill virus, pathogens of red grouse. Journal of Animal Ecology, 72, 177–185.CrossRefGoogle Scholar
Law, R. (2000). Fishing, selection, and phenotypic evolution. ICES Journal of Marine Science, 57, 659–668.CrossRefGoogle Scholar
Law, R. and Dicekmann, U. (1998). Symbiosis through exploitation and the merger of lineages in evolution. Proceedings of the Royal Society of London, B265, 1245–1253.CrossRefGoogle Scholar
Law, R., Murrell, D. J. and Dieckmann, U. (2003). Population growth in space and time: spatial logistic equations. Ecology, 84, 252–262.CrossRefGoogle Scholar
Leggett, W. C. and Deblois, E. (1994). Recruitment in marine fishes: is it regulated by starvation and predation in the egg and larval stages?Netherlands Journal of Sea Research, 32, 119–134.CrossRefGoogle Scholar
Lenski, R. E. and May, R. M. (1994). The evolution of virulence in parasites and pathogens: reconciliation between two competing hypotheses. Journal of Theoretical Biology, 169, 253–265.CrossRefGoogle ScholarPubMed
Leonard, T. and Hsu, J. S. J. (1999). Bayesian Methods. Cambridge: Cambridge University Press.Google Scholar
Levin, S. A. and Segel, L. A. (1985). Pattern generation in space and aspect. SIAM Review, 27, 45–67.CrossRefGoogle Scholar
Levins, R. (1966). The strategy of model building in population biology. American Scientist, 54, 421–431.Google Scholar
Levins, R. (1969). Some demographic and genetic consequences of environmental heterogeneity for biological control. Bulletin of the Entomological Society of America, 15, 237–240.CrossRefGoogle Scholar
Levins, R. (1970). Extinction. In Gerstenhaber, M. (editor) Some Mathematical Questions in Biology. Providence, RI: American Mathematical Society, pp. 75–107.Google Scholar
Lewin, R. (1986). Punctuated equilibrium is now old hat. Science, 231, 672–673.CrossRefGoogle ScholarPubMed
Lewis, W. J., and Takasu, K. (1990). Use of learned odours by a parasitic wasp in accordance with host and food needs. Nature, 348, 635–636.CrossRefGoogle Scholar
Lewontin, R. and Cohen, D. (1969). On population growth in a randomly varying environment. Proceedings of the National Academy of Sciences, 62, 1056–1060.CrossRefGoogle Scholar
Liermann, M. and Hilborn, R. (1997). Depensation in fish stocks: a hierarchic Bayesian meta-analysis. Canadian Journal of Fisheries and Aquatic Sciences, 54, 1976–1984.CrossRefGoogle Scholar
Lighthill, M. J. (1958). Introduction to Fourier Analysis and Generalised Functions. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lima, S. L. (2002). Putting predators back into behavioral predator-prey interactions. Trends in Ecology and Evolution, 17, 70–75.CrossRefGoogle Scholar
Lin, C. C. and Segel, L. A. (1988 (1974)). Mathematics Applied to Deterministic Problems in the Natural Sciences. Philadelphia: SIAM (Society for Industrial and Applied Mathematics).CrossRefGoogle Scholar
Lin, Z.-S. (2003). Simulating unintended effects restoration. Ecological Modelling, 164, 169–175.CrossRefGoogle Scholar
Lindholm, J. B., Auster, P. J., Ruth, M. and Kaufman, L. (2001). Modeling the effects of fishing and implications for the design of marine protected areas: juvenile fish responses to variations in seafloor habitat. Conservation Biology, 15, 424–437.CrossRefGoogle Scholar
Lindley, S. T. (2003). Estimation of population growth and extinction parameters from noisy data. Ecological Applications, 13, 806–813.CrossRefGoogle Scholar
Link, D. R., Natale, G., Shao, R., et al. (1997). Spontaneous formation of macroscopic chiral domains in a fluid smectic phase of achiral molecules. Science, 278, 1924–1927.CrossRefGoogle Scholar
Link, J. S., Brodziak, J. K. T., Edwards, S. F., et al. (2002). Marine ecosystem assessment in a fisheries management context. Canadian Journal of Fisheries and Aquatic Sciences, 59, 1429–1440.CrossRefGoogle Scholar
Lipp, E. K., Huq, A. and Colwell, R. K. (2002). Effects of global climate on infectious disease: the cholera model. Clinical Microbiology Reviews, 15, 757–770.CrossRefGoogle ScholarPubMed
Lochmiller, R. L. and Deerenberg, C. (2000). Trade-offs in evolutionary immunology: just what is the cost of immunity?Oikos, 88, 87–98.CrossRefGoogle Scholar
Lockwood, D. R., Hastings, A. M. and Botsford, L. W. (2002). The effects of dispersal patterns on marine reserves: does the tail wag the dog?Theoretical Population Biology, 61, 297–310.CrossRefGoogle ScholarPubMed
LoGiudice, K., Ostfeld, R. S., Schmidt, K. A. and Keesing, F. (2003). The ecology of infectious disease: effects of host diversity and community composition on Lyme disease risk. Proceedings of the National Academy of Science, 100, 567–571.CrossRefGoogle ScholarPubMed
Lorenzen, K. (2000). Population dynamics and management. In Beveridge, M. C. M. and McAndrew, B. J. (editors) Tilapias: Biology and Exploitation. Dordrecht: Kluwer Academic Publishers, pp. 163–226.CrossRefGoogle Scholar
Ludwig, D. (1975). Persistence of dynamical systems under random perturbations. SIAM Review, 17, 605–640.CrossRefGoogle Scholar
Ludwig, D. (1981). Escape from domains of attraction for systems perturbed by noise. In Enns, R. H., Jones, B. L., Miura, R. M. and Rangnekar, S. S. (editors) Nonlinear Phenomena in Physics and Biology. New York: Plenum Press.CrossRefGoogle Scholar
Ludwig, D. (1995). Uncertainty and fisheries management. Lecture Notes in Biomathematics, 100, 516–528.CrossRefGoogle Scholar
Ludwig, D. (1999). Is it meaningful to estimate a probability of extinction?Ecology, 80, 298–310.CrossRefGoogle Scholar
Ludwig, D., Mangel, M. and Haddad, B. (2001). Ecology, conservation and public policy. Annual Review of Ecology and Systematics, 32, 481–517.CrossRefGoogle Scholar
MacArthur, R. H. and Pianka, E. R. (1966). On the optimal use of a patchy environment. American Naturalist, 100, 603–609.CrossRefGoogle Scholar
MacArthur, R. H. and Wilson, E. O. (1967). The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
MacCall, A. D. (1990). Dynamic Geography of Marine Fish Populations. Seattle, WA: University of Washington Press.Google Scholar
MacCall, A. D. (1998). Use of decision tables to develop a precautionary approach to problems in behavior, life history and recruitment variability. In Restrepo, V. R. (editor) Proceedings of the Fifth NMFS Stock Assessment Workshop: Providing Scientific Advice to Implement the Precautionary Approach under the Magnuson–Stevens Fishery Conservation and Management Act. Key Largo, FL: US Department of Commerce.Google Scholar
MacCall, A. D. (2002). Use of known-biomass production models to determine productivity of West Coast groundfish stocks. North American Journal of Fisheries Management, 22, 272–279.2.0.CO;2>CrossRefGoogle Scholar
MacDonald, N. (1989). Biological Delay Systems: Linear Stability Theory. Cambridge: Cambridge University Press.Google Scholar
Mackauer, M. and Voelkl, W. (1993). Regulation of aphid populations by aphidiid wasps: does parasitoid foraging behaviour or hyperparasitism limit impact? Oecologia, 94, 339–350.CrossRefGoogle ScholarPubMed
Malloch, P. D. (1994). Life-History and Habits of the Salmon, Sea-Trout, Trout and Other Freshwater Fish. Derrydale Press.Google Scholar
Mangel, M. (1982). Applied mathematicians and naval operators. SIAM Review, 24, 289–300.CrossRefGoogle Scholar
Mangel, M. (1985). Decision and Control in Uncertain Resource Systems. New York: Academic Press.Google Scholar
Mangel, M. (1992). Descriptions of superparasitism by optimal foraging theory, evolutionary stable strategies, and quantitative genetics. Evolutionary Ecology, 6, 152–169.CrossRefGoogle Scholar
Mangel, M. (1998). No-take areas for sustainability of harvested species and a conservation invariant for marine reserves. Ecology Letters, 1, 87–90.CrossRefGoogle Scholar
Mangel, M. (2000a). Irreducible uncertainties, sustainable fisheries and marine reserves. Evolutionary Ecology Research, 2, 547–557.Google Scholar
Mangel, M. (2000b). On the fraction of habitat allocated to marine reserves. Ecology Letters, 3, 15–22.CrossRefGoogle Scholar
Mangel, M. (2000c). Trade-offs between fish habitat and fishing mortality and the role of reserves. Bulletin of Marine Science, 66, 663–674.Google Scholar
Mangel, M. (2001a). Complex adaptive systems, aging and longevity. Journal of Theoretical Biology, 213, 559–571.CrossRefGoogle Scholar
Mangel, M. (2001b). Required reading for (ecological) battles. (Review of N Eldredge. The Triumph of Evolution … and the Failure of Creationism). Trends in Ecology and Evolution, 16, 110.CrossRefGoogle Scholar
Mangel, M. and Beder, J. H. (1985). Search and stock depletion: theory and applications. Canadian Journal of Fisheries and Aquatic Sciences, 42, 150–163.CrossRefGoogle Scholar
Mangel, M. and Clark, C. W. (1988). Dynamic Modeling in Behavioral Ecology. Princeton, NJ: Princeton University Press.Google Scholar
Mangel, M. and Ludwig, D. (1977). Probability of extinction in a stochastic competition. SIAM Journal on Applied Mathematics, 33, 256–266.CrossRefGoogle Scholar
Mangel, M. and Roitberg, B. D. (1992). Behavioral stabilization of host–parasitoid population dynamics. Theoretical Population Biology, 42, 308–320.CrossRefGoogle Scholar
Mangel, M. and Tier, C. (1993). Dynamics of metapopulations with demographic stochasticity and environmental catastrophes. Theoretical Population Biology, 44, 1–31.CrossRefGoogle Scholar
Mangel, M. and Tier, C. (1994). Four facts every conservation biologist should know about persistence. Ecology, 75, 607–614.CrossRefGoogle Scholar
Mangel, M., Hofman, R. J., Norse, E. A. and Twiss, J. R. (1993). Sustainability and ecological research. Ecological Applications, 3, 573–575.Google ScholarPubMed
Mangel, M., Mullan, A., Mulch, A., Staub, S. and Yasukochi, E. (1998). A generally accessible derivation of the golden rule of bioeconomics. Bulletin of the Ecological Society of America, 79, 145–148.Google Scholar
Mangel, M., Fiksen, O. and Giske, J. (2001). Theoretical and statistical models in natural resource management and research. In Shenk, T. M. and Franklin, A. B. (editors) Modeling in Natural Resource Management. Development, Interpretation, and Application. Washington, DC: Island Press, pp. 57–72.Google Scholar
Mantua, N. J., Hare, S. R., Zhang, Y., Wallace, J. M. and Francis, R. C. (1997). A Pacific interdecadal climate oscillation with impacts on salmon production. Bulletin of the American Meteorological Society, 78, 1069–1079.2.0.CO;2>CrossRefGoogle Scholar
Margulis, L. and Sagan, D. (2002). Acquiring Genomes. A Theory of the Origins of Species. New York: Basic Books.Google Scholar
Marris, G., Hubbar, S. and Hughes, J. (1986). Use of patchy resources by Nemeritis canescens (Hymenoptera: Ichneumonidae). I. Optimal solutions. Journal of Animal Ecology, 55, 631–640.CrossRefGoogle Scholar
Marshall, C. T., Kjesbu, O. S., Yaragina, N. A., Solemdal, P. and Ulltang, O. (1998). Is spawner biomass a sensitive measure of the reproductive and recruitment potential of Northeast Arctic cod? Canadian Journal of Fisheries and Aquatic Sciences, 55, 1766–1783.CrossRefGoogle Scholar
Martz, H. F. and Waller, R. A. (1982). Bayesian Reliability Analysis. New York: John Wiley and Sons.Google Scholar
Masel, J. and Bergman, A. (2003). The evolution of the evolvability properties of the yeast prion ‘PSI + ’. Evolution, 57, 1498–1512.CrossRefGoogle Scholar
Matsuda, H. and Nishimori, K. (2003). A size-structured model for a stock-recovery program for an exploited endemic fisheries resource. Fisheries Research, 60, 223–236.CrossRefGoogle Scholar
Maunder, M. (2003). Is it time to discard the Schaefer model from the stock assessment scientist's toolbox? Fisheries Research, 61, 145–149.CrossRefGoogle Scholar
Maunder, M. N. (2002). The relationship between fishing methods, fisheries management and the estimation of maximum sustainable yield. Fish and Fisheries, 3, 251–260.CrossRefGoogle Scholar
May, R. M. (1974). Stability and Complexity in Model Ecosystems, 2nd edn. Princeton, NJ: Princeton University Press.Google Scholar
May, R. M. and Anderson, R. M. (1978). Regulation and stability of host-parasite population interactions. II. Destabilizing processes. Journal of Animal Ecology, 47, 249–267.CrossRefGoogle Scholar
May, R. M., Conway, G. R., Hassel, M. P. and Southwood, T. R. E. (1974). Time delays, density dependence and single species oscillations. Journal of Animal Ecology, 43, 747–770.CrossRefGoogle Scholar
Maynard Smith, J. (1968). Some Mathematical Ideas in Biology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Maynard Smith, J. (1982). Evolution and the Theory of Games. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
McAllister, M. K. (1996). Applications of Bayesian decision theory to fisheries policy formation: review. In Arnason, R. and Davidsson, T. B. (editors) Essays on Statistical Modelling Methodology for Fisheries Management. Rekjavek, Iceland: The Fisheries Research Institute University Press.Google Scholar
McAllister, M. K. and Ianelli, J. N. (1997). Bayesian stock assessment using catch-age data and the sampling-importance resampling algorithm. Canadian Journal of Fisheries and Aquatic Sciences, 54, 284–300.Google Scholar
McAllister, M. K. and Kirkwood, G. P. (1998a). Bayesian stock assessment: a review and example application using the logistic model. ICES Journal of Marine Science, 55, 1031–1060.CrossRefGoogle Scholar
McAllister, M. K. and Kirkwood, G. P. (1998b). Using Bayesian decision analysis to help achieve a precautionary approach for managing developing fisheries. Canadian Journal of Fisheries and Aquatic Sciences, 55, 2642–2661.CrossRefGoogle Scholar
McAllister, M. K. and Kirkwood, G. P. (1999). Applying multivariate conjugate priors in fishery-management system evaluation: how much quicker is it and does it bias the ranking of management options. ICES Journal of Marine Science, 56, 884.CrossRefGoogle Scholar
McAllister, M. K. and Peterman, R. M. (1992a). Decision analysis of a large-scale fishing experiment designed to test for a genetic effect of size-selective fishing on British Columbia pink salmon (Oncorhynchus gorbuscha). Canadian Journal of Fisheries and Aquatic Sciences, 49, 1305–1314.CrossRefGoogle Scholar
McAllister, M. K. and Peterman, R. M. (1992b). Experimental design in the management of fisheries: A review. North American Journal of Fisheries Management, 12, 1–18.2.3.CO;2>CrossRefGoogle Scholar
McAllister, M. K., Pikitch, E. K., Punt, A. E. and Hilborn, R. (1994). A Bayesian approach to stock assessment and harvest decision using the sampling/importance resampling algorithm. Canadian Journal of Fisheries and Aquatic Sciences, 51, 2673–2687.CrossRefGoogle Scholar
McAllister, M. K., Starr, P. J., Restrepo, V. R. and Kirkwood, G. P. (1999). Formulating quantitative methods to evaluate fishery-management systems: what fishery processes should be modelled and what trade-offs should be made? ICES Journal of Marine Science, 56, 900–916.CrossRefGoogle Scholar
McAllister, M. K., Pikitch, E. K. and Babcock, E. A. (2001). Using demographic methods to construct Bayesian priors for the intrinsic rate of increase in the Schaefer model and implications for stock rebuilding. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1871–1890.CrossRefGoogle Scholar
McCallum, H., Barlow, N. and Hone, J. (2001). How should pathogen transmission be modelled? Trends in Ecology and Evolution, 16, 295–300.CrossRefGoogle ScholarPubMed
McClanahan, T. R. and Kaunda-Arara, B. (1996). Fishery recovery in a coral-reef marine park and its effect on the adjacent fishery. Conservation Biology, 10, 1187–1199.CrossRefGoogle Scholar
McDonald, A. D., Sandal, L. K. and Steinshamn, S. I. (2002). Implications of a nested stochastic/deterministic bio-economic model for a pelagic fishery. Ecological Modelling, 149, 193–201.CrossRefGoogle Scholar
McDowall, R. M. (1988). Diadromy in Fishes. Migrations Between Freshwater and Marine Environments. Portland, OR: Timber Press.Google Scholar
McGarvey, R. (2003). Demand-side fishery management: integrating two forms of input control. Marine Policy, 27, 207–218.CrossRefGoogle Scholar
McGregor, R. (1997). Host-feeding and oviposition by parasitoids on hosts of different fitness value: influences of egg load and encounter rate. Journal of Insect Behavior, 10, 451–462.CrossRefGoogle Scholar
McGregor, R. R. and Roitberg, B. D. (2000). Size-selective oviposition by parasitoids and the evolution of life-history timing in hosts: fixed preference vs frequency-dependent host selection. Oikos, 89, 305–312.CrossRefGoogle Scholar
Mchich, R., Auger, P. M., Parra, R. B. and Raissi, N. (2002). Dynamics of a fishery on two fishing zones with fish stock dependent migrations: aggregation and control. Ecological Modelling, 158, 51–62.CrossRefGoogle Scholar
McLeod, P., Martin, A. P. and Richards, K. J. (2002). Minimum length scale for growth-limited oceanic plankton distributions. Ecological Modelling, 158, 111–120.CrossRefGoogle Scholar
McNamara, J. M., Houston, A. I. and Collins, E. J. (2001). Optimality models in behavioral biology. SIAM Review, 43, 413–466.CrossRefGoogle Scholar
McNeill, J. R. (2000). Something New Under the Sun: an Environmental History of the Twentieth-Century World. New York: Norton.Google Scholar
McPhee, J. (2002). The Founding Fish. New York: Farrar, Straus and Giroux.Google Scholar
McPhee, M. V. and Quinn, T. P. (1998). Factors affecting the duration of nest defense and reproductive lifespan of female sockeye salmon, Oncorhynchus nerka. Environmental Biology of Fishes, 51, 469–475.Google Scholar
Medvinsky, A. B., Petrovskii, S. V., Tikhonova, I. A., Malchow, H. and Li, B.-L. (2002). Spatiotemporal complexity of plankton and fish dynamics. SIAM Review, 44, 311–370.CrossRefGoogle Scholar
Meier, C., Senn, W., Hauser, R. and Zimmerman, M. (1994). Strange limits of stability in host-parasitoid systems. Journal of Mathematical Biology, 32, 563–572.CrossRefGoogle Scholar
Meltzer, M. (2003). Risks and benefits of preexposure and postexposure smallpox vaccination. Emerging Infectious Diseases, 9, 1363–1370.CrossRefGoogle ScholarPubMed
Meltzer, M. I., Damon, I., LeDuc, J. W. and Millar, D. J. (2001). Modeling potential responses to smallpox as a bioterrorist weapon. Emerging Infectious Diseases, 7, 959–968.CrossRefGoogle ScholarPubMed
Merton, R. C. (1971). Optimum consumption and portfolio rules in a continuous-time model. Journal of Economic Theory, 3, 373–413.CrossRefGoogle Scholar
Mesnil, B. (2003). The catch-survey analysis (CSA) method of fish stock assessment: an evaluation using simulated data. Fisheries Research, 63, 193–212.CrossRefGoogle Scholar
Meyer, R. and Millar, R. B. (1999a). Bayesian stock assessment using a state-space implementation of the delay difference model. Canadian Journal of Fisheries and Aquatic Sciences, 56, 37–52.Google Scholar
Meyer, R. and Millar, R. B. (1999b). BUGS in Bayesian stock assessments. Canadian Journal of Fisheries and Aquatic Sciences, 56, 1078–1086.CrossRefGoogle Scholar
Millar, R. B. (2002). Reference priors for Bayesian fisheries models. Canadian Journal of Fisheries and Aquatic Sciences, 59, 1492–1502.CrossRefGoogle Scholar
Millar, R. B. and Meyer, R. (2000). Bayesian state-space modeling of age-structured data: fitting a model is just the beginning. Canadian Journal of Fisheries and Aquatic Sciences, 57, 43–50.CrossRefGoogle Scholar
Mills, D. (1989). Ecology and Management of Atlantic Salmon. London: Chapman and Hall.Google Scholar
Mitchell, W. A. and Valone, T. J. (1990). The Optimization Research Program: studying adaptations by their function. The Quarterly Review of Biology, 65, 43–52.CrossRefGoogle Scholar
Moerland, T. S. (1995). Temperature: enzyme and organelle. In Hochackha, P. and Mommsen, T. P. (editors) Biochemistry and Molecular Biology of Fishes. Amsterdam: Elsevier Science B.V.Google Scholar
Moret, Y. and Schmid-Hempel, P. (2000). Survival for immunity: the price of immune system activation for bumblebee workers. Science, 290, 1166–1167.CrossRefGoogle Scholar
Morgan, E. R., Milner-Gulland, E. J., Torgerson, P. R. and Medley, G. F. (2004). Ruminating on complexity: macroparasites of wildlife and livestock. Trends in Ecology and Evolution, 19, 181–188.CrossRefGoogle ScholarPubMed
Morris, W. F., Mangel, M. and Adler, F. R. (1995). Mechanisms of pollen deposition by insect pollinators. Evolutionary Ecology, 9, 304–317.CrossRefGoogle Scholar
Morse, P. M. and Kimball, G. E. (1951). Methods of Operations Research. Technology Press of Massachusetts Institute of Technology; and Cambridge and New York: Wiley.Google Scholar
Mosmann, T. R. and Sad, S. (1996). The expanding universe of T-cell subsets: Th1, Th2 and more. Immunology Today, 17, 138–146.CrossRefGoogle ScholarPubMed
Mosquera, J. and Adler, F. R. (1998). Evolution of virulence: a unified framework for coinfection and superinfection. Journal of Theoretical Biology, 195, 293–313.CrossRefGoogle ScholarPubMed
Mosquera, J., Côté, M., Jennings, S. and Reynolds, J. D. (2000). Conservation benefits of marine reserves for fish populations. Animal Conservation, 4, 321–332.CrossRefGoogle Scholar
Mueller, L. D. and Rose, M. R. (1996). Evolutionary theory predicts late-life mortality plateaus. Proceedings of the National Academy of Sciences, 93, 15 249–15 253.CrossRefGoogle ScholarPubMed
Mullan, M. (1993). Webs and Scales. Physical and Ecological Processes in Marine Recruitment. Seattle, WA: University of Washington Press.Google Scholar
Murawski, S. A. (2000). Definitions of overfishing from an ecosystem perspective. ICES Journal of Marine Science, 57, 649–658.CrossRefGoogle Scholar
Murdoch, W. W. (1994). Population regulation in theory and practice. Ecology, 75, 271–287.CrossRefGoogle Scholar
Murdoch, W. W., Nisbet, R., Blythe, S. P., Gurney, W. S. C. and Reeve, J. D. (1987). An invulnerable age class and stability in delay-differential parasitoid–host models. American Naturalist, 129, 263–282.CrossRefGoogle Scholar
Murdoch, W. W., Briggs, C. J. and Nisbet, R. (2003). Consumer–Resource Dynamics. Princeton, NJ: Princeton University Press.Google Scholar
Murray, J. D. (1990). Turing's theory of morphogenesis – its influence on modelling biological pattern and form. Bulletin of Mathematical Biology, 52, 119–153.CrossRefGoogle Scholar
Murray, J. D. (2002). Mathematical Biology I: An Introduction. New York: Springer Verlag.Google Scholar
Murray, J. D. (2003). Mathematical Biology II: Spatial Models and Biomedical Applications. New York: Springer Verlag.Google Scholar
Myers, R. A., Mertz, G. and Fowlow, P. S. (1997a). Maximum population growth rates and recovery times for Atlantic cod, Gadus morhua. Fishery Bulletin, 95, 762–772.Google Scholar
Myers, R. A., Hutchings, J. A. and Barrowman, N. J. (1997b). Why do fish stocks collapse? The example of cod in Atlantic Canada. Ecological Applications, 7, 91–106.CrossRefGoogle Scholar
Myers, R. A., MacKenzie, B. R., Bowen, K. G. and Barrowman, N. J. (2001). What is the carrying capacity for fish in the ocean? A meta-analysis of population dynamics of North Atlantic cod. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1464–1476.CrossRefGoogle Scholar
Myers, R. A., Barrowman, N. J., Hilborn, R. and Kehler, D. G. (2002). Inferring Bayesian priors with limited direct data: Applications to risk analysis. North American Journal of Fisheries Management, 22, 351–364.2.0.CO;2>CrossRefGoogle Scholar
Nakken, O., Sandberg, P. and Steinshamn, S. I. (1996). Reference points for optimal fish stock management. Marine Policy, 20, 447–462.CrossRefGoogle Scholar
Narasimhan, T. N. (1999). Fourier's heat conduction equation: history, influence, and connections. Reviews of Geophysics, 37, 151–172.CrossRefGoogle Scholar
Nasell, I. (2002). Stochastic models of some endemic infections. Mathematical Biosciences, 179, 1–19.CrossRefGoogle ScholarPubMed
Needle, C. L. (2002). Recruitment models: diagnosis and prognosis. Reviews in Fish Biology and Fisheries, 11, 95–111.CrossRefGoogle Scholar
Nesse, R. M. and Williams, G. C. (1994). Why We Get Sick. New York: Vintage Books.Google Scholar
Nicholson, A. J. (1933). The balance of animal populations. Journal of Animal Ecology, 2, 131–178.CrossRefGoogle Scholar
Nicholson, A. J. (1954). An outline of the dynamics of animal populations. Australian Journal of Zoology, 2, 9–65.CrossRefGoogle Scholar
Nicholson, A. J. and Bailey, V. A. (1935). The balance of animal populations. Proceedings of the Zoological Society of London, 3, 551–598.CrossRefGoogle Scholar
Nisbet, R. and Gurney, W. S. C. (1982). Modelling Fluctuating Populations. New York: John Wiley and Sons.Google Scholar
Nisbet, R. M. and Gurney, W. S. C. (1983). The systematic formulation of population models for insects with dynamically varying instar duration. Theoretical Population Biology, 23, 114–135.CrossRefGoogle Scholar
Nisbet, R. M., Blythe, S. P., Gurney, W. S. C. and Metz, J. A. J. (1985). Stage-structure models of populations with distinct growth and development processes. IMA Journal of Mathematics Applied in Medicine and Biology, 2, 57–68.CrossRefGoogle ScholarPubMed
Norse, E. A. (editor) (1993). Global Marine Biological Diversity. Washington, DC: Island Press.Google Scholar
Nowak, M. A. and May, R. M. (2000). Virus Dynamics. Mathematical Principles of Immunology and Virology. Oxford: Oxford University Press.Google Scholar
Nuland, S. B. (1993). How We Die. New York: Vintage Books.Google Scholar
O'Neill, P. D. (2002). A tutorial introduction to Bayesian inference for stochastic epidemic models using Markov chain Monte Carlo methods. Mathematical Biosciences, 180, 103–114.CrossRefGoogle ScholarPubMed
Oaks, S. C. Jr., Mitchell, V. S., Pearson, G. W. and Carpenter, C. C. J. (editors) (1991). Malaria. Obstacles and Opportunities. Washington, DC: National Academy Press.Google Scholar
Okey, T. A. (2003). Membership of the eight Regional Fishery Management Councils in the United States: are special interests over-represented? Marine Policy, 27, 193–206.CrossRefGoogle Scholar
Olson, D. M. and Andow, D. A. (1997). Primary sex allocation in Trichogramma (Hymenoptera: Trichogrammatidae) and the effects of sperm on oviposition behavior. Annals of the Entomological Society of America, 90, 689–692.CrossRefGoogle Scholar
Olson, D. M., Fadamiro, H., Lundgren, J. G. and Heimpel, G. E. (2000). Effects of sugar feeding on carbohydrate and lipid metabolism in a parasitoid wasp. Physiological Entomology, 25, 17–26.CrossRefGoogle Scholar
Olver, C., Shuter, B. J. and Minns, C. K. (1995). Toward a definition of conservation principles for fisheries management. Canadian Journal of Fisheries and Aquatic Sciences, 52, 1584–1594.CrossRefGoogle Scholar
Overholtz, W. J. (1999). Precision and uses of biological reference points calculated from stock recruitment data. North American Journal of Fisheries Management, 19, 643–657.2.0.CO;2>CrossRefGoogle Scholar
Overholtz, W. J., Edwards, S. F. and Brodziak, J. K. T. (1995). Effort control in the New England groundfish fishery: a bioeconomic perspective. Canadian Journal of Fisheries and Aquatic Sciences, 52, 1944–1957.CrossRefGoogle Scholar
Owen-Smith, N. (2002). Adaptive Herbivore Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Pacala, S. W., Hassell, M. P. and May, R. M. (1990). Host–parasitoid associations in patchy environments. Nature, 344, 150–153.CrossRefGoogle ScholarPubMed
Packer, C., Holt, R., Hudson, P., Lafferty, K. and Dobson, A. (2003). Keeping the herds healthy and alert: implications of predator control for infectious disease. Ecology Letters, 6, 797–802.CrossRefGoogle Scholar
Paddack, M. J. and Estes, J. A. (2000). Kelp forest fish populations in marine reserves and adjacent exploited areas of Central California. Ecological Applications, 10, 855–870.CrossRefGoogle Scholar
Pandolfini, B. (1989). Chess Openings: Traps and Zaps. New York: Simon and Schuster.Google Scholar
Patterson, K., Cook, R., Darby, C., et al. (2001). Estimating uncertainty in fish stock assessment and forecasting. Fish and Fisheries, 2, 125–157.CrossRefGoogle Scholar
Patterson, K. R. (1998). Assessing fish stocks when catches are misreported: model, simulation tests, and application to cod, haddock and whiting in the ICES area. ICES Journal of Marine Science, 55, 878–891.CrossRefGoogle Scholar
Patterson, K. R. (1999). Evaluating uncertainty in harvest control law catches using Bayesian Markov chain Monte Carlo virtual population analysis with adaptive rejection sampling and including structural uncertainty. Canadian Journal of Fisheries and Aquatic Sciences, 56, 208–221.CrossRefGoogle Scholar
Pearcy, W. G. (1992). Ocean Ecology of Pacific Salmonids. Seattle, WA: University of Washington Press.Google Scholar
Pearl, R. (1928). The Rate of Living. New York: Alfred Knopf.Google Scholar
Pearl, R. and Miner, J. R. (1935). Experimental studies on the duration of life. XIV. The comparative mortality of certain lower organisms. Quarterly Review of Biology, 10, 60–79.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1921). Experimental studies on the duration of life. I. Introductory discussion of the duration of life in drosophila. American Naturalist, 55, 481–509.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1922a). Experimental studies on the duration of life. II. Hereditary differences in duration of life in the line-bred strains of drosophila. American Naturalist, 56, 174–187.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1922b). Experimental studies on the duration of life. III. The effect of successive etherizations on the duration of life in drosophila. American Naturalist, 56, 273–280.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1922c). Experimental studies on the duration of life. IV. Data on the influence of density of population on duration of life in drosophila. American Naturalist, 56, 312–321.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1922d). Experimental studies on the duration of life. V. On the influence of certain environmental factors on the duration of life in drosophila. American Naturalist, 56, 385–405.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1924a). Experimental studies on the duration of life. IX. New life tables for drosophila. American Naturalist, 58, 71–82.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1924b). Experimental studies on the duration of life. X. The duration of life of drosophila melanogaster in the complete absence of food. American Naturalist, 58, 193–218.CrossRefGoogle Scholar
Pearl, R., Miner, J. R. and Parker, S. A. (1927). Experimental studies on the duration of life. XI. Density of population and life duration in drosophila. American Naturalist, 56, 289–318.CrossRefGoogle Scholar
Pearl, R., Parker, S. A. and Gonzalez, B. M. (1923). Experimental studies on the duration of life. VII. The Mendelian inheritance of duration of life in crosses of wild type and quintuple stocks of drosophila melanogaster. American Naturalist, 57, 153–192.CrossRefGoogle Scholar
Pearl, R., Parker, T. and Miner, J. R. (1941). Experimental studies on the duration of life. XVI. Life tables for the flour beetle Tribolium confusum duval. American Naturalist, 75, 5–19.CrossRefGoogle Scholar
Pellmyr, O. (2003). Yuccas, yucca moths, and coevolution: a review. Annals of the Missouri Botanical Garden, 90, 35–55.CrossRefGoogle Scholar
Perelson, A. S. and Nelson, P. W. (1999). Mathematical analysis of HIV-1 dynamics in vivo. SIAM Review, 41, 3–44.CrossRefGoogle Scholar
Perkins, P. C. and Edwards, E. F. (1995). A mixture model for estimating discarded bycatch from data with many zero observations: tuna discards in the eastern tropical Pacific Ocean. Fishery Bulletin, 94, 330–340.Google Scholar
Petchey, O. L. (2000). Environmental colour affects aspects of single-species population dynamics. Proceedings of the Royal Society of London, B267, 747–754.CrossRefGoogle ScholarPubMed
Petchey, O. L., Gonzalez, A. and Wilson, H. B. (1997). Effects on population persistence: the interaction between environmental noise colour, intraspecific competition and space. Proceedings of the Royal Society of London, B264, 1841–1847.CrossRefGoogle Scholar
Peterman, R. M. (1989). Application of statistical power analysis to the Oregon coho salmon (Oncorhynchus kisutch) problem. Canadian Journal of Fisheries and Aquatic Sciences, 46, 1183–1187.CrossRefGoogle Scholar
Peterman, R. M. (1990a). The importance of reporting statistical power: the forest decline and acid deposition example. Ecology, 71, 2024–2027.CrossRefGoogle Scholar
Peterman, R. M. (1990b). Statistical power analysis can improve fisheries research and management. Canadian Journal of Fisheries and Aquatic Sciences, 47, 2–15.CrossRefGoogle Scholar
Peters, C. S., Mangel, M. and Costantino, R. F. (1989). Stationary distribution of population size in Tribolium. Bulletin of Mathematical Biology, 51, 625–638.Google Scholar
Peters, R. H. (1983). The Ecological Implications of Body Size. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Peters, R. H. (1991). A Critique for Ecology. Cambridge: Cambridge University Press.Google Scholar
Pezzey, J. C. V., Roberts, C. M. and Urdal, B. T. (2000). A simple bioeconomic model of a marine reserve. Ecological Economics, 33, 77–91.CrossRefGoogle Scholar
Pielke, R. A. Jr. and Conant, R. T. (2003). Best practices in prediction for decision-making: lessons from the atmospheric and earth sciences. Ecology, 84, 1351–1358.CrossRefGoogle Scholar
Pierce, N. E. and Nash, D. R. (1999). The Imperial Blue: Jalmenus evagoras (Lycaenidae). In Kitching, R. L., Jones, R. E., Pierce, N. E., and Scheermeyer, E., eds., Biology of Australian Butterflies. Victoria: CSIRO Publishing, pp. 277–313.Google Scholar
Pierce, N. E., Braby, M. F., Heath, A., et al. (2002). The ecology and evolution of ant association in the Lycaenidae (Lepidoptera). Annual Review of Entomology, 47, 733–771.CrossRefGoogle Scholar
Pigliucci, M. and Murren, C. J. (2003). Genetical assimilation and a possible evolutionary paradox: can macroevolution sometimes be so fast as to pass us by? Evolution, 57, 1455–1464.CrossRefGoogle Scholar
Pikitch, E. K., Huppert, D. D. and Sissenwine, M. P. (editors) (1997). Global Trends: Fishery Management. Bethesda, MD: American Fisheries Society.Google Scholar
Pikitch, E. K., Santora, C., Babcock, E. A., et al. (2004). Ecosystem-based fishery management. Science, 305, 346–347.CrossRefGoogle ScholarPubMed
Pincock, R. E. and Wilson, K. R. (1973). Spontaneous generation of optical activity. Journal of Chemical Education, 50, 455–457.CrossRefGoogle Scholar
Pincock, R. E., Perkins, R. R., Ma, A. S. and Wilson, K. R. (1971). Probability distribution of enantiomorphous forms in spontaneous generation of optically active substances. Science, 174, 1018–1020.CrossRefGoogle ScholarPubMed
Pirsig, R. M. (1974). Zen and the Art of Motorcycle Maintenance. New York: William Morrow.Google Scholar
Pitcher, T. J. (2000). Ecosystem goals can reinvigorate fisheries management, help dispute resolution and encourage public support. Fish and Fisheries, 1, 99–103.CrossRefGoogle Scholar
Pitcher, T. J., Watson, R., Forrest, R., Valtysson, H. P. and Guénette, S. (2002). Estimating illegal and unreported catches from marine ecosystems: a basis for change. Fish and Fisheries, 3, 317–339.CrossRefGoogle Scholar
Pletcher, S. D. and Curtsinger, J. W. (1998). Mortality plateaus and the evolution of senescence: why are old-age mortality rates so low? Evolution, 52, 454–464.CrossRefGoogle ScholarPubMed
Prager, M. H. (2002). Comparison of logistic and generalized surplus-production models applied to swordfish, Xiphias gladius, in the north Atlantic Ocean. Fisheries Research, 58, 41–57.CrossRefGoogle Scholar
Prager, M. H., Porch, C. E., Shertzer, K. W. and Caddy, J. F. (2003). Targets and limits for management of fisheries: a simple probability-based approach. North American Journal of Fisheries Management, 23, 349–361.2.0.CO;2>CrossRefGoogle Scholar
Preston, S. H., Heuveline, P. and Guillot, M. (2001). Demography. Measuring and Modeling Population Processes. Oxford: Blackwell.Google Scholar
Pritchard, G. (1969). The ecology of a natural population of Queensland fruit fly, Dacus tryoni. Australian Journal of Zoology, 17, 293–311.CrossRefGoogle Scholar
Pritchett, V. S. (1990a). The Complete Collected Stories. New York: Random House.Google Scholar
Pritchett, V. S. (1990b). The Complete Collected Essays. New York: Random House.Google Scholar
Provine, W. B. (1986). Sewall Wright and Evolutionary Biology. Chicago, IL: University of Chicago Press.Google Scholar
Punt, A. E. and Hilborn, R. (1997). Fisheries stock assessment and decision analysis: the Bayesian approach. Reviews in Fish Biology and Fisheries, 7, 35–63.CrossRefGoogle Scholar
Puterman, M. L. (1994). Markov Decision Processes. Discrete Stochastic Dynamic Programming. New York: John Wiley and Sons.Google Scholar
Pybus, O., Charleston, M. A., Gupta, S., et al. (2001). The epidemic behavior of the hepatitis C virus. Science, 292, 2323–2325.CrossRefGoogle ScholarPubMed
Quinn, T. J. I. and Deriso, R. B. (1999). Quantitative Fish Dynamics. New York: Oxford University Press.Google Scholar
Raftery, A. E. (1988). Inference for the binomial N-parameter: A hierarchical Bayes approach. Biometrika, 75, 223–228.CrossRefGoogle Scholar
Railsback, S. F. (2001). Concepts from complex adaptive systems as a framework for individual-based modelling. Ecological Modelling, 139, 47–62.CrossRefGoogle Scholar
Read, A. F. (2003). Simplicity and serenity in advanced statistics. Trends in Ecology and Evolution, 18, 11–12.CrossRefGoogle Scholar
Read, A. F., Anwar, M., Shutler, D. and Nee, S. (1995). Sex allocation and population structure in malaria and related parasitic protozoa. Proceedings of the Royal Society of London, B260, 359–363.CrossRefGoogle ScholarPubMed
Read, A. F., Narara, A., Nee, S., Keymer, A. E. and Day, K. P. (1992). Gametocyte sex ratios as indirect measures of outcrossing rates in malaria. Parasitology, 104, 387–395.CrossRefGoogle ScholarPubMed
Reed, W. J. and Simons, C. M. (1996). Analyzing catch-effort data by means of the Kalman filter. Canadian Journal of Fisheries and Aquatic Sciences, 53, 2157–2166.Google Scholar
Rees, M., Mangel, M., Turnbull, L., Sheppard, A. and Briese, D. (2000). The effects of heterogeneity on dispersal and colonization in plants. In Hutchings, M. J., John, E. A., and Stewart, A. J. A., eds., The Ecological Consequences of Environmental Heterogeneity. Oxford: Blackwell Science, pp. 237–265.Google Scholar
Reid, C. (1976). Courant in Göttingen and New York. The Story of an Improbable Mathematician. New York: Springer Verlag.Google Scholar
Reidl, J. and Klose, K. E. (2002). Vibrio cholerae and cholera: out of the water and into the host. FEMS Microbiology Reviews, 26, 125–139.CrossRefGoogle Scholar
Ricciardi, L. M. and Sato, S. (1990). Diffusion processes and first-passage-time problems. In Ricciardi, L. M., ed., Lectures in Applied Mathematics and Informatics. New York: St. Martin's Press, pp. 206–284.Google Scholar
Richards, L. J. and Maguire, J.-J. (1998). Recent international agreements and the precautionary approach: new directions for fisheries management science. Canadian Journal of Fisheries and Aquatic Sciences, 55, 1545–1552.CrossRefGoogle Scholar
Rickman, S. J., Dulvy, N. K., Jennings, S. and Reynolds, J. D. (2000). Recruitment variation related to fecundity in marine fishes. Canadian Journal of Fisheries and Aquatic Sciences, 57, 116–124.CrossRefGoogle Scholar
Rivot, E., Prevost, E. and Parent, E. (2001). How robust are Bayesian posterior inferences based on a Ricker model with regards to measurement errors and prior assumptions about parameters? Canadian Journal of Fisheries and Aquatic Sciences, 58, 2284–2297.CrossRefGoogle Scholar
Robb, C. A. and Peterman, R. M. (1998). Application of Bayesian decision analysis to management of a sockeye salmon (Oncorhynchus nerka) fishery. Canadian Journal of Fisheries and Aquatic Sciences, 55, 86–98.CrossRefGoogle Scholar
Robertson, I. C., Robertson, W. G. and Roitberg, B. D. (1998). A model of mutual tolerance and the origin of communal associations between unrelated females. Journal of Insect Behavior, 11, 265–286.CrossRefGoogle Scholar
Roff, D. A. (1984). The evolution of life history parameters in teleosts. Canadian Journal of Fisheries and Aquatic Sciences, 41, 989–999.CrossRefGoogle Scholar
Roff, D. A. (1991). The evolution of life-history variation in fishes, with particular reference to flatfishes. Netherlands Journal of Sea Research, 27, 197–207.CrossRefGoogle Scholar
Rohani, P., Keeling, M. J. and Grenfell, B. T. (2002). The interplay between determinism and stochasticity in childhood diseases. American Naturalist, 159, 469–481.CrossRefGoogle ScholarPubMed
Roitberg, B. D. (1990). Optimistic and pessimistic fruit flies: evaluating fitness consequences of estimation errors. Behaviour, 114, 65–82.CrossRefGoogle Scholar
Roitberg, B. D. and Friend, W. G. (1992). A general theory for host seeking decisions in mosquitoes. Bulletin of Mathematical Biology, 54, 401–412.CrossRefGoogle ScholarPubMed
Roitberg, B. D. and Lalonde, R. G. (1991). Host marking enhance parasitism risk for a fruit-infesting fly Rhagoletis basiola. Oikos, 61, 389–393.CrossRefGoogle Scholar
Roitberg, B. D. and Prokopy, R. J. (1987). Insects that mark hosts. BioScience, 37, 400–406.CrossRefGoogle Scholar
Roitberg, B. D., Cailrl, R. S. and Prokopy, R. J. (1984). Oviposition deterring pheromone influences dispersal distance in tephritid fruit flies. Entomologia Experimentalis et Applicata, 35, 217–220.CrossRefGoogle Scholar
Roitberg, B. D., Mangel, M., Lalonde, R., et al. (1992). Seasonal dynamic shifts in patch exploitation by parasitic wasps. Behavioral Ecology, 3, 156–165.CrossRefGoogle Scholar
Roitberg, B. D., Sircom, J., Roitberg, C. A., Alphen, J. J. M. and Mangel, M. (1993). Life expectancy and reproduction. Nature, 364, 351.CrossRefGoogle ScholarPubMed
Rolff, J. and Siva-Jothy, M. T. (2002). Copulation corrupts immunity: a mechanism for a cost of mating in insects. PNAS, 99, 9916–9918.CrossRefGoogle ScholarPubMed
Rolff, J. and Siva-Jothy, M. T. (2003). Invertebrate ecological immunology. Science, 301, 472–475.CrossRefGoogle ScholarPubMed
Romagnani, S. (1996). TH1 and TH2 in human diseases. Clinical Immunology and Immunopathology, 80, 225–235.CrossRefGoogle ScholarPubMed
Root, K. V. (1998). Evaluating the effects of habitat quality, connectivity, and catastrophes on threatened species. Ecological Applications, 8, 854–865.CrossRefGoogle Scholar
Rosà, R., Pugliese, A., Villani, A. and Rizzoli, A. (2003). Individual-based vs. deterministic models for macroparasites: host cycles and extinction. Theoretical Population Biology, 63, 295–307.CrossRefGoogle ScholarPubMed
Rosenberg, A. A. (2003). Managing to the margins: the overexploitation of fisheries. Frontiers in Ecology and the Environment, 1, 102–106.CrossRefGoogle Scholar
Rosenberg, A. A. and Restrepo, V. R. (1994). Uncertainty and risk evaluation in stock assessment advice for U.S. marine fisheries. Canadian Journal of Fisheries and Aquatic Sciences, 51, 2715–2720.CrossRefGoogle Scholar
Rosenheim, J. A. (1996). An evolutionary argument for egg limitation. Evolution, 50, 2089–2094.CrossRefGoogle ScholarPubMed
Rosenheim, J. A. (1999). Characterizing the cost of oviposition in insects: a dynamic model. Evolutionary Ecology, 13, 141–165.CrossRefGoogle Scholar
Rosenheim, J. A. and Heimpel, G. E. (1998). Egg limitation in parasitoids: A review of the evidence and a case study. Biological Control, 11, 160–168.Google Scholar
Rosenheim, J. A. and Heimpel, G. E. (2000). Egg maturation, egg resorption, and the costliness of transient egg limitation in insects. Proceedings of the Royal Society of London, B267, 1565–1573.CrossRefGoogle ScholarPubMed
Rosenheim, J. A. and Mangel, M. (1994). Patch-leaving rules for parasitoids with imperfect host discrimination. Ecological Entomology, 19, 374–380.CrossRefGoogle Scholar
Rosenheim, J. A. and Rosen, D. (1992). Influence of egg load and host size on host-feeding behaviour of the parasitoid Aphytis lingnaensis. Ecological Entomology, 17, 263–272.CrossRefGoogle Scholar
Rosenzweig, M. L. (1995). Species Diversity in Space and Time. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Ross, R. (1916). An application of the theory of probabilities to the study of a priori pathometry. Part I. Proceedings of the Royal Society of London, A92, 204–230.CrossRefGoogle Scholar
Ross, R. and Hudson, H. P. (1917a). An application of the theory of probabilities to the study of a priori pathometry. Part II. Proceedings of the Royal Society of London, A93, 212–225.CrossRefGoogle Scholar
Ross, R. and Hudson, H. P. (1917b). An application of the theory of probabilities to the study of a priori pathometry. Part III. Proceedings of the Royal Society of London, A93, 225–240.CrossRefGoogle Scholar
Royall, R. (1997). Statistical Evidence. A Likelihood Paradigm. London: Chapman and Hall.Google Scholar
Rudd, M. A., Tupper, M. H., Folmer, H. and Kooten, G. C. (2003). Policy analysis for tropical marine reserves: challenges and directions. Fish and Fisheries, 4, 65–85.CrossRefGoogle Scholar
Ruxton, G. D. and Rohani, P. (1996). The consequences of stochasticity for self-organized spatial dynamics, persistence and coexistence in spatially extended host-parasitoid communities. Proceedings of the Royal Society of London, B263, 625–631.CrossRefGoogle Scholar
Sadovy, Y. (2001). The threat of fishing to highly fecund fishes. Journal of Fish Biology, 59 (Supplement A), 90–108.CrossRefGoogle Scholar
Saether, B.-E., Engen, S., Lande, R., Both, C. and Visser, M. E. (2002). Density dependence and stochastic variation in a newly established population of a small songbird. Oikos, 99, 331–337.CrossRefGoogle Scholar
Saether, B. E. and Engen, S. (2004). Stochastic population theory faces reality in the laboratory. Trends in Ecology and Evolution, 19, 351–353.CrossRefGoogle ScholarPubMed
Sanchez Lizaso, J. L., Goni, R., Renones, O., et al. (2000). Density dependence in marine protected populations: a review. Environmental Conservation, 27, 144–158.CrossRefGoogle Scholar
Sanchirico, J. N., Stoffle, R., Broad, K. and Talaue-McManus, L. (2003). Modeling marine protected areas. Science, 301, 47–48.CrossRefGoogle ScholarPubMed
Sauvage, F., Langlais, M., Yoccoz, N. G. and Pontier, D. (2003). Modelling hantavirus in fluctuating populations of bank voles: the role of indirect transmission on virus persistence. Journal of Animal Ecology, 72, 1–13.CrossRefGoogle Scholar
Scheffer, M. and Carpenter, S. (2003). Catastrophic regime shifts in ecosystems: linking theory to observation. Trends in Ecology and Evolution, 18, 648–656.CrossRefGoogle Scholar
Scheffer, M., Carpenter, S., Foley, J. A., Folke, C. and Walker, B. (2001). Catastrophic shifts in ecosystems. Nature, 413, 591–596.CrossRefGoogle ScholarPubMed
Schemske, D. W. and Bierzychudek, P. (2001). Perspective: evolution of flower color in the desert annual Linanthus parryae: Wright revisited. Evolution, 55, 1269–1282.CrossRefGoogle ScholarPubMed
Schluter, D. (2000). The Ecology of Adaptive Radiation. New York: Oxford University Press.Google Scholar
Schmid-Hempel, P. and Ebert, D. (2003). On the evolutionary ecology of specific immune defense. Trends in Ecology and Evolution, 18, 27–32.CrossRefGoogle Scholar
Schnute, J., Cass, A. and Richards, L. J. (2000). A Bayesian decision analysis to set escapement goals for Fraser River sockeye salmon (Oncorhynchus nerka). Canadian Journal of Fisheries and Aquatic Sciences, 57, 962–979.CrossRefGoogle Scholar
Schnute, J. T. (1993). Ambiguous inferences from fisheries data. In Barnett, V. and Turkman, K. F. (editors) Statistics for the Environment. London: John Wiley and Sons, pp. 293–309.Google Scholar
Schnute, J. T. and Kronlund, A. R. (1996). A management oriented approach to stock recruitment analysis. Canadian Journal of Fisheries and Aquatic Sciences, 53, 1281–1293.CrossRefGoogle Scholar
Schnute, J. T. and Kronlund, A. R. (2002). Estimating salmon stock-recruitment relationships from catch and escapement data. Canadian Journal of Fisheries and Aquatic Sciences, 59, 433–449.CrossRefGoogle Scholar
Schnute, J. and Richards, L. J. (1998). Analytical models for fishery reference points. Canadian Journal of Fisheries and Aquatic Sciences, 55, 515–528.CrossRefGoogle Scholar
Schnute, J. T. and Richards, L. J. (2001). Use and abuse of fishery models. Canadian Journal of Fisheries and Aquatic Sciences, 58, 10–17.CrossRefGoogle Scholar
Schrag, S. J. and Wiener, A. P. (1995). Emerging infectious disease: what are the relative roles of ecology and evolution? Trends in Ecology and Evolution, 10, 319–324.CrossRefGoogle ScholarPubMed
Schulman, L. S. (1981). Techniques and Applications of Path Integration. New York: Wiley Interscience.Google Scholar
Schuss, Z. (1980). Theory and Applications of Stochastic Differential Equations. New York: Wiley.Google Scholar
Schwartz, M. (2003). How the Cows Turned Mad. Berkeley, CA: University of California Press.Google Scholar
Schwarz, G. (1978). Estimating the dimension of a model. Annals of Statistics, 6, 461–464.CrossRefGoogle Scholar
Seber, G. A. F. (1982). The Estimation of Animal Abundance, 2nd edn. New York: MacMillan.Google Scholar
Settle, W. H. and Wilson, L. T. (1990). Invasion by the variegated leafhopper and biotic interactions: parasitism, competition, and apparent competition. Ecology, 71, 1461–1470.CrossRefGoogle Scholar
Sevenster, J. A., Ellers, J. and Driessen, G. (1998). An evolutionary argument for time limitation. Evolution, 52, 1241–1244.CrossRefGoogle ScholarPubMed
Shaw, D. J. and Dobson, A. P. (1995). Patterns of macroparasite abundance and aggregation in wildlife populations: a quantitative review. Parasitology, 111, S111–S113.CrossRefGoogle ScholarPubMed
Shears, N. T. and Babcock, R. C. (2003). Continuing trophic cascade effects after 25 years of no-take marine reserve protection. Marine Ecology – Progress Series, 246, 1–16.CrossRefGoogle Scholar
Sheldon, B. C. and Verhulst, S. (1996). Ecological immunology: costly parasite defences and trade-offs in evolutionary ecology. TREE, 11, 317–321.Google ScholarPubMed
Shelton, P. A. and Healey, B. P. (1999). Should depensation be dismissed as a possible explanation for the lack of recovery of the northern cod (Gadus morhua) stock? Canadian Journal of Fisheries and Aquatic Sciences, 56, 1521–1524.CrossRefGoogle Scholar
Shepherd, J. G. (1982). A versatile new stock recruitment relationship for fisheries, and the construction of sustainable yield curves. J. Conseil International Exploration de la Mer, 40, 67–75.CrossRefGoogle Scholar
Shepherd, J. G. (1999). Extended survivors analysis: An improved method for the analysis of catch-at-age data and abundance indices. ICES Journal of Marine Science, 56, 584–591.CrossRefGoogle Scholar
Sherman, K. and Alexander, L. M. (editors) (1989). Biomass Yields and Geography of Large Marine Ecosystems. Boulder, CO: Westview Press.Google Scholar
Sherman, K., Alexander, L. M. and Gold, B. D. (editors) (1990). Large Marine Ecosystems. Patterns, Process and Yields. Washington, DC: American Association for the Advancement of Science.Google Scholar
Sherman, K., Alexander, L. M. and Gold, B. D. (editors) (1991). Food Chains, Yields, Models and Management of Large Marine Ecosystems. Boulder, CO: Westview Press.Google Scholar
Sherman, K., Alexander, L. M. and Gold, B. D. (editors) (1993). Large Marine Ecosystems. Stress, Mitigation and Sustainability. Washington, DC: American Association for the Advancement of Science.Google Scholar
Shudo, E. and Iwasa, Y. (2001). Inducible defense against pathogens and parasites: optimal choice among multiple options. Journal of Theoretical Biology, 209, 233–247.CrossRefGoogle ScholarPubMed
Siegel, J. S. (2002). Shattered mirrors. Nature, 419, 346–347.CrossRefGoogle ScholarPubMed
Simberloff, D. (1980). A succession of paradigms in ecology: essentialism to materialism and probabilism. Synthese, 43, 3–39.CrossRefGoogle Scholar
Sinclair, A. F. and Swain, D. P. (1996). Comment: spatial implications of a temperature-based growth model for Atlantic Cod (Gadus morhua) off the eastern coast of Canada. Canadian Journal of Fisheries and Aquatic Sciences, 53, 2909–2911.CrossRefGoogle Scholar
Sinclair, A. F., Swain, D. P. and Hanson, J. M. (2002a). Disentangling the effects of size-selective mortality, density, and temperature on length-at-age. Canadian Journal of Fisheries and Aquatic Sciences, 59, 372–382.CrossRefGoogle Scholar
Sinclair, M., Arnason, R., Csirke, J., et al. (2002b). Responsible fisheries in the marine ecosystem. Fisheries Research, 58, 255–265.CrossRefGoogle Scholar
Singleton, D. A. and Vo, L. K. (2002). Enantioselective synthesis without discrete optically active additives. Journal of the American Chemical Society, 124, 10 010–10 011.CrossRefGoogle ScholarPubMed
Siwoff, S., Hirdt, S., Hirdt, P. and Hirdt, T. (1990). The 1990 Elias Baseball Analyst. New York: Collier Books.Google Scholar
Skellam, J. G. (1951). Random dispersal in theoretical populations. Biometrika, 38, 196–218.CrossRefGoogle ScholarPubMed
Skorokhod, A. V., Hoppensteadt, F. C. and Salehi, H. (2002). Random Perturbation Methods. New York: Springer Verlag.CrossRefGoogle Scholar
Slepoy, A., Singh, R. R. P., Pázmándi, F., Kulkarni, R. V. and Cox, D. L. (2001). Statistical mechanics of prion diseases. Physical Review Letters, 87, 058101.CrossRefGoogle ScholarPubMed
Smith, G. C. and Wilkinson, D. (2002). Modelling disease spread in a novel host: rabies in the European badger Meles meles. Journal of Applied Ecology, 39, 865–874.CrossRefGoogle Scholar
Smith, M. C., Reeder, R. H. and Thomas, M. B. (1997). A model to determine the potential for biological control of Rottboellia cochinchinensis with the head smut Sporisorium ophiuri. Journal of Applied Ecology, 34, 388–398.CrossRefGoogle Scholar
Smith, R. D. (2000). Can't You Hear Me Callin'. The Life of Bill Monroe. Boston: Little, Brown and Company.Google Scholar
Smith, T. D. (1994). Scaling Fisheries. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Snow, C. P. (1965). The Two Cultures. Cambridge: Cambridge University Press.Google Scholar
Soai, K., Niwa, S. and Hori, H. (1990). Asymmetric self-catalytic reaction. Self-production of chiral 1-(3-pyridyl)alkanols as chiral self-catalysts in the enantioselective addition of dialkylzinc reagents to pyridine-3-caraldehyde. Journal of the Chemical Society, Chemical Communications, 1990, 982–983.CrossRefGoogle Scholar
Soh, S., Gunderson, D. R. and Ito, D. H. (2000). The potential role of marine reserves in the management of shortraker rockfish (Sebastest borealis) and rougheye rockfish (S. aleutianus) in the Gulf of Alaska. Fishery Bulletin, 99, 168–179.Google Scholar
Soule, M. E. (editor) (1987). Viable Populations for Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Sounes, H. (2001). Down the Highway. The Life of Bob Dylan. New York: Grove Press.Google Scholar
Southwood, T. R. E. (1978). Ecological Methods. London: Chapman and Hall.Google Scholar
Souza, P. D. (1998). On hyperbolic discounting and uncertain hazard rates. Proceedings of the Royal Society of London, B265, 2015–2020.CrossRefGoogle Scholar
Spielman, A. and D'Antonio, M. (2001). Mosquito. New York: Hyperion.Google Scholar
Stachel, J. (editor) (1998). Einstein's Miraculous Year. Five Papers that Changed the Face of Physics. Princeton, NJ: Princeton University Press.Google Scholar
Starr, P., Annala, J. H. and Hilborn, R. (1998). Contested stock assessment: two case studies. Canadian Journal of Fisheries and Aquatic Sciences, 55, 529–537.CrossRefGoogle Scholar
Stearns, S. C. (editor) (1999). Evolution in Health and Disease. New York: Oxford University Press.Google Scholar
Stearns, S. C. (2000). Daniel Bernoulli (1738): evolution and economics under risk. Journal of Biosciences, 25, 221–228.CrossRefGoogle Scholar
Stenseth, N. C., Bjørnstad, O. N., Falck, W., et al. (1999). Dynamics of coastal cod populations: intra- and intercohort density dependence and stochastic processes. Proceedings of the Royal Society of London, A266, 1645–1654.CrossRefGoogle Scholar
Stephens, D. W. and Krebs, J. R. (1986). Foraging Theory. Princeton, NJ: Princeton University Press.Google Scholar
Stewart, I. (2000). The Lorenz attractor exists. Nature, 406, 948–949.CrossRefGoogle ScholarPubMed
Stockwell, C. A., Hendry, A. P., Berg, O. K., Quinn, T. P. and Kinnison, M. T. (2003). Contemporary evolution meets conservation biology. Trends in Ecology and Evolution, 18, 94–101.CrossRefGoogle Scholar
Stolz, J. and Schnell, J. (editors) (1991). Trout. Harrisburg, PA: Stackpole Books.Google Scholar
Strand, E., Huse, G. and Giske, J. (2002). Artificial evolution of life history and behavior. American Naturalist, 159, 624–644.CrossRefGoogle ScholarPubMed
Stratonovich, R. L. (1963). Topics in the Theory of Random Noise. New York: Gordon and Breach.Google Scholar
Stratoudakis, Y., Fryer, R. J. and Cook, R. M. (1998). Discarding practices for commercial gadoids in the North sea. Canadian Journal of Fisheries and Aquatic Sciences, 55, 1632–1644.CrossRefGoogle Scholar
Strogatz, S. H. (1994). Nonlinear Dynamics and Chaos. Cambridge, MA: Westview.Google Scholar
Stroock, D. W. (2003). Markov Processes from K. Ito's Perspective. Princeton, NJ: Princeton University Press.CrossRefGoogle Scholar
Strunk, W. J. and White, E. B. (1979). The Elements of Style, 3rd edn. New York: MacMillan.Google Scholar
Sumpter, D. J. T. and Martin, J. (2004). The dynamics of virus epidemics in Varroa-infested honey bee colonies. Journal of Animal Ecology, 73, 51–63.CrossRefGoogle Scholar
Suter, G. W. (1996). Abuse of hypothesis testing statistics in ecological risk assessment. Human and Ecological Risk Assessment, 2, 331–347.CrossRefGoogle Scholar
Swansburg, E., Chaput, G., Moore, D., Caissie, D. and El-Jabi, N. (2002). Size variability of juvenile Atlantic salmon: links to environmental conditions. Journal of Fish Biology, 61, 661–683.CrossRefGoogle Scholar
Sylvain, G., Baalen, M. and Jansen, V. A. A. (2002). The evolution of parasite virulence, superinfection, and host resistance. American Naturalist, 159, 658–669.Google Scholar
Tanaka, M. M., Kumm, J. and Feldman, M. W. (2002). Coevolution of pathogens and cultural practices: a new look at behavioral heterogeneity in epidemics. Theoretical Population Biology, 62, 111–119.CrossRefGoogle Scholar
Tang, S. and Wang, Y. (2002). A parameter estimation program for the error-in-variable model. Ecological Modelling, 156, 225–236.CrossRefGoogle Scholar
Taper, M. L. and Lele, S. R. (2004). The Nature of Scientific Evidence. Statistical, Philosophical and Empirical Considerations. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Taylor, A. D. (1988a). Large-scale spatial structure and population dynamics in arthropod predator–prey systems. Annals Zoologia Fennici, 25, 63–74.Google Scholar
Taylor, A. D. (1988b). Parasitoid competition and the dynamics of host–parasitoid models. American Naturalist, 132, 417–436.CrossRefGoogle Scholar
Taylor, A. D. (1993). Aggregation, competition, and host–parasitoid dynamics: stability conditions don't tell it all. American Naturalist, 141, 501–506.CrossRefGoogle Scholar
Taylor, L. R. (1984). Assessing and interpreting the spatial distributions of insect populations. Annual Review of Entomology, 29, 321–357.CrossRefGoogle Scholar
Taylor, L. R., Woiwood, I. P. and Perry, J. N. (1978). The density-dependence of spatial behaviour and the rarity of randomness. Journal of Animal Ecology, 47, 383–406.CrossRefGoogle Scholar
Taylor, L. R., Woiwood, I. P. and Perry, J. N. (1979). The negative binomial as a dynamic ecological model for aggregation, and the density dependence of k. Journal of Animal Ecology, 48, 289–304.CrossRefGoogle Scholar
Taylor, L. R., Woiwood, I. P. and Perry, J. N. (1980). Variance and the large scale spatial stability of aphids, moths and birds. Journal of Animal Ecology, 49, 831–854.CrossRefGoogle Scholar
Haar, D. (1998). Master of Modern Physics. The Scientific Contributions of H. A. Kramers. Princeton, NJ: Princeton University Press.Google Scholar
Thom, R. (1972/1975). Stabilité structurelle et morphogenese (Structural Stability and Morphogenesis). Reading, MA: Benjamin.Google Scholar
Thompson, G. G. (1992). A Bayesian approach to management advice when stock-recruitment parameters are uncertain. Fishery Bulletin, 90, 561–573.Google Scholar
Thompson, G. G. (1993). A proposal for a threshold stock size and maximum fishing mortality rate. In Smith, S. J., Hunt, J. J. and Rivard, D. (editors) Risk Evaluation and Biological Reference Points for Fisheries Management (Canadian Special Publication of Fisheries and Aquatic Sciences 120). Ottawa: National Research Council of Canada.Google Scholar
Toft, C. A. and Mangel, M. (1991). From individuals to ecosystems: the papers of Skellam, Lindeman and Hutchinson. Bulletin of Mathematical Biology, 53, 121–134.CrossRefGoogle Scholar
Trumble, R. J. (1998). Northeast Pacific flatfish management. Journal of Sea Research, 39, 167–181.CrossRefGoogle Scholar
Turchin, P. (1998). Quantitative Analysis of Movement. Measuring and Modeling Population Redistribution in Animals and Plans. Sunderland, MA: Sinauer Associates.Google Scholar
Turelli, M. A., Schemske, D. W. and Bierzychudek, P. (2001). Stable two-allele polymorphisms maintained by fluctuating fitnesses and seed banks: protecting the blues in Linanthus parryae. Evolution, 55, 1283–1298.CrossRefGoogle ScholarPubMed
Turing, A. M. (1952). The chemical basis of morphogenesis. Proceedings of the Royal Society of London, B237, 37–72.CrossRefGoogle Scholar
Tyutyunov, Y., Senina, I., Jost, C. and Arditi, R. (2002). Risk assessment of the harvested pike-perch population of the Azov Sea. Ecological Modelling, 149, 297–311.CrossRefGoogle Scholar
Uhlenbeck, G. E. and Ornstein, L. S. (1930). On the theory of Brownian motion. Physical Review, 36, 823–841.CrossRefGoogle Scholar
Ulrich, C., Gallic, B., Dunn, M. R. and Gascuel, D. (2002a). A multi-species multi-fleet bioeconomic simulation model for the English Channel artisinal fisheries. Fisheries Research, 58, 379–401.CrossRefGoogle Scholar
Ulrich, C. and Marchal, P. (2002). Sensitivity of some biological reference points to shifts in exploitation patterns and inputs uncertainty for three North Sea demersal stocks. Fisheries Research, 58, 153–169.CrossRefGoogle Scholar
Ulrich, C., Pascoe, D., Sparre, P. J., Wilde, J.-W. and Marchal, P. (2002b). Influence of trends in fishing power on bioeconomics in the North Sea flatfish fishery regulated by catches or by effort quotas. Canadian Journal of Fisheries and Aquatic Sciences, 59, 829–843.CrossRefGoogle Scholar
Valleron, A.-J., Boelle, P.-Y., Will, R. and Cesbron, J.-Y. (2001). Estimation of epidemic size and incubation time based on age characteristics of vCJD in the United Kingdom. Science, 294, 1726–1728.CrossRefGoogle ScholarPubMed
Alphen, J. J. M. and Visser, M. E. (1990). Superparasitism as an adaptive strategy. Annual Review of Entomology, 35, 59–79.CrossRefGoogle ScholarPubMed
Alphen, J. J. M., Dijken, M. J. and Waage, J. K. (1987). A functional approach to superparasitism: host discrimination needs not to be learnt. Netherlands Journal of Zoology, 37, 167–179.CrossRefGoogle Scholar
Alphen, J. J. M., Bernstein, C. and Driessen, G. (2003). Information acquisition and time allocation in insect parasitoids. Trends in Ecology and Evolution, 18, 81–87.CrossRefGoogle Scholar
van Baalen, M. (2000). The evolution of parasitoid egg load. In Hochberg, M. E. and Ives, A. R. (editors). Parasitoid Population Biology. Princeton, NJ: Princeton University Press.Google Scholar
Kampen, N. G. (1977). A soluble model for diffusion in a bistable potential. Journal of Statistical Physics, 17, 71–88.CrossRefGoogle Scholar
Kampen, N. G. (1981a). Ito versus Stratonovich. Journal of Statistical Physics, 24, 175–187.CrossRefGoogle Scholar
Kampen, N. G. (1981b). Stochastic Processes in Physics and Chemistry. Amsterdam: North-Holland Publishing Company.Google Scholar
Oostenbrugge, J. A. E., Densen, W. L. T. and Machiels, M. A. M. (2001). Risk aversion in allocating fishing effort in a highly uncertain coastal fishery for pelagic fish, Moluccas, Indonesia. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1683–1691.CrossRefGoogle Scholar
Randen, E. J. and Roitberg, B. D. (1996). The effect of egg load on superparasitism by the snowberry fly. Entomologia Experimentalis et Applicata, 79, 241–245.CrossRefGoogle Scholar
Winkle, W., Jager, H. I., Railsback, S. F., et al. (1998). Individual-based model of sympatric populations of brown and rainbow trout for instream flow assessment: model description and calibration. Ecological Modelling, 110, 175–207.CrossRefGoogle Scholar
Vasseur, D. A. and Yodzis, P. (2004). The color of environmental noise. Ecology, 85, 1146–1152.CrossRefGoogle Scholar
Vaupel, J. W., Carey, J. R., Christensen, K., et al. (1998). Biodemographic trajectories of longevity. Science, 280, 855–859.CrossRefGoogle ScholarPubMed
Vestergaard, N. (1996). Discard behavior, highgrading and regulation: the case of the Greenland Shrimp Fishery. Marine Resource Economics, 11, 247–266.CrossRefGoogle Scholar
Vestergaard, N., Squires, D. and Kirkley, J. (2003). Measuring capacity and capacity utilization in fisheries: the case of the Danish Gill-net fleet. Fisheries Research, 60, 357–368.CrossRefGoogle Scholar
Visser, M. E. and Rosenheim, J. A. (1998). The influence of competition between foragers on clutch size decisions in insect parasitoids. Biological Control, 11, 169–174.CrossRefGoogle Scholar
Bertalanffy, L. (1957). Quantitative laws in metabolism and growth. Quarterly Review of Biology, 32, 217–231.CrossRefGoogle Scholar
Voronka, R. and Keller, J. (1975). Asymptotic analysis of stochastic models in population genetics. Mathematical Biosciences, 25, 331–362.CrossRefGoogle Scholar
Vos, M. and Hemerik, L. (2003). Linking foraging behavior to lifetime reproductive success for an insect parasitoid: adaptation to host distributions. Behavioral Ecology, 14, 236–245.CrossRefGoogle Scholar
Vos, M., Hemerik, L. and Vet, L. E. M. (1998). Patch exploitation by the parasitoids Cotesia rubecula and Cotesia glomerata in multi-patch environments with different host distributions. Journal of Animal Ecology, 67, 774–783.CrossRefGoogle Scholar
Wachter, K. W. (1999). Evolutionary demographic models for mortality plateaus. Proceedings of the National Academy of Sciences, 96, 10 544–10 547.CrossRefGoogle ScholarPubMed
Wachter, K. W. and Finch, C. E. (editors) (1997). Between Zeus and the Salmon. The Biodemography of Longevity. Washington, DC: National Academy Press.Google Scholar
Wade, P. R. (2000). Bayesian methods in conservation biology. Conservation Biology, 14, 1308–1316.CrossRefGoogle Scholar
Waeckers, F. L. (1994). The effect of food deprivation on the innate visual and olfactory preferences in the parasitoid Cotesia rubecula. Journal of Insect Physiology, 40, 641–649.CrossRefGoogle Scholar
Wajnberg, E., Fauveergue, X. and Pons, O. (2000). Patch leaving decision rules and the marginal value theorem: an experimental analysis and a simulation model. Behavioral Ecology, 11, 577–586.CrossRefGoogle Scholar
Walde, S. J. and Murdoch, W. W. (1988). Spatial density dependence in parasitoids. Annual Review of Entomology, 33, 441–466.CrossRefGoogle Scholar
Walter, G. H. (1988). Activity patterns and egg production in Coccophagus bartletti, an aphelinid parasitoid of scale insects. Ecological Entomology, 13, 95–105.CrossRefGoogle Scholar
Walters, C. J. and Martell, S. J. D. (2004). Fisheries Ecology and Management. Princeton, NJ: Princeton University Press.Google Scholar
Waser, P. M., Elliott, L. F., Creel, N. M. and Creel, S. R. (1995). Habitat variation and viverrid demography. In Sinclair, A. R. E. and Arcese, P. (editors) Serengeti II: Dynamics, Management, and Conservation of an Ecosystem. Chicago, IL: University of Chicago Press, pp. 421–447.Google Scholar
Washburn, A. R. (1981). Search and Detection. Arlington, VA: Military Applications Section, Operations Research Society of America.Google Scholar
Watson, R. (1993). The Trout. A Fisherman's Natural History. Shrewsbury, UK: Swan Hill Press.Google Scholar
Watson, R. (1999). Salmon, Trout and Charr of the World. A Fisherman's Natural History. Shrewsbury, UK: Swan Hill Press.Google Scholar
Wegner, K., Kalbe, M., Kurtz, J., Reusch, T. and Milinski, M. (2003). Parasite selection for immunogenetic optimality. Science, 301, 1343.CrossRefGoogle ScholarPubMed
Weitz, J. S. and Fraser, H. B. (2001). Explaining mortality plateaus. Proceedings of the National Academy of Sciences, 98, 15 383–15 386.CrossRefGoogle Scholar
West, G. B., Brown, J. H. and Enquist, B. J. (1997). A general model for the origin of allometric scaling laws in biology. Science, 276, 122–126.CrossRefGoogle ScholarPubMed
West, G. B., Brown, J. H. and Enquist, B. J. (2001). A general model for ontogenetic growth. Nature, 413, 628–631.CrossRefGoogle ScholarPubMed
West, M. and Harrison, J. (1997). Bayesian Forecasting and Dynamic Models. New York: Springer Verlag.Google Scholar
West-Eberhard, M. J. (2003). Developmental Plasticity and Evolution. New York: Oxford University Press.Google Scholar
Wheatley, D. N. and Agutter, P. S. (1996). Historical aspects of the origin of diffusion theory in 19th-century mechanistic materialism. Perspectives in Biology and Medicine, 40, 139–155.CrossRefGoogle Scholar
White, A., Bowers, R. G. and Begon, M. (1996). Red/blue chaotic power spectra. Nature, 381, 198.CrossRefGoogle ScholarPubMed
White, P. C. L. and Harris, S. (1995). Bovine tuberculosis in badger (Meles meles) populations in southwest England: the use of a spatial stochastic simulation model to understand the dynamics of the disease. Philosophical Transactions of the Royal Society of London, B349, 391–413.CrossRefGoogle ScholarPubMed
Whittle, P. (1955). The outcome of a stochastic epidemic. A note on Bailey's paper. Biometrika, 42, 116–122.Google Scholar
Whittle, P. (1983). Optimization over Time. Dynamic Programming and Stochastic Control. Volume II. New York: John Wiley and Sons.Google Scholar
Wilcox, C. and Elderd, B. (2003). The effect of density-dependent catastrophes on population persistence time. Journal of Applied Ecology, 40, 859–871.CrossRefGoogle Scholar
Wilcox, C. and Possingham, H. (2002). Do life history traits affect the accuracy of diffusion approximations for mean time to extinction? Ecological Applications, 12, 1163–1179.CrossRefGoogle Scholar
Wilkinson, D. and Sherratt, T. N. (2001). Horizontally acquired mutualisms, an unsolved problem in ecology? Oikos, 92, 377–384.CrossRefGoogle Scholar
Williams, E. H. (2002). The effects of unaccounted discards and misspecified natural mortality on harvest policies based on estimates of spawners per recruit. North American Journal of Fisheries Management, 22, 311–325.2.0.CO;2>CrossRefGoogle Scholar
Williams, G. C. and Nesse, R. M. (1991). The dawn of Darwinian medicine. Quarterly Review of Biology, 66, 1–22.CrossRefGoogle ScholarPubMed
Williams, J. G. (2003). Sardine fishing in the early 20th century. Science, 300, 2032–2033.CrossRefGoogle ScholarPubMed
Wilmott, P. (1998). Derivatives. The Theory and Practice of Financial Engineering. Chichester: John Wiley and Sons.Google Scholar
Wilson, W. G., Morris, W. F. and Bronstein, J. L. (2003). Coexistence of mutualists and exploiters on spatial landscapes. Ecological Monographs, 73, 397–413.CrossRefGoogle Scholar
Wong, E. (1971). Stochastic Processes in Information and Dynamical Systems. New York: McGraw-Hill.Google Scholar
Wong, E. and Zakai, M. (1965). On the relation between ordinary and stochastic differential equations. International Journal of Engineering Science, 3, 213–229.Google Scholar
Wooster, W. S. (editor) (1988). Fishery Science and Management. Objectives and Limitations. New York: Springer Verlag.CrossRefGoogle Scholar
Wouk, H. (1962). Youngblood Hawke. New York: Doubleday.Google Scholar
Wouk, H. (1994). The Glory. New York: Little, Brown.Google Scholar
Wynia, M. K. and Gostin, L. (2002). The bioterrorist threat and access to health care. Science, 296, 1613.CrossRefGoogle ScholarPubMed
Yates, F. (1951). The influence of Statistical Methods for Research Workers on the development of the science of statistics. Journal of the American Statistical Association, 46, 19–34.Google Scholar
Yildiz, F. H. and Schoolnik, G. K. (1999). Vibrio cholerae O1 El Tor: Identification of a gene cluster required for the rugose colony type, exopolysaccharide production, chlorine resistance, and biofilm formation. Proceedings of the National Academy of Sciences, 96, 4028–4033.CrossRefGoogle ScholarPubMed
Young, T. P. (1994). Natural die-offs of large mammals: implications for conservation. Conservation Biology, 8, 410–418.CrossRefGoogle Scholar
Zabel, R. W., Harvey, C. J., Katz, S. L., Good, T. P. and Levin, P. S. (2003). Ecologically sustainable yield. American Scientist, 91, 150–157.Google Scholar
Zinnser, W. (1989). Writing to Learn. New York: Harper and Row.Google Scholar
Zinsser, W. (2001). On Writing Well. New York: Harper Collins.Google Scholar
Zuk, M. and Stoehr, A. M. (2002). Immune defense and host life history. American Naturalist, 160, S9–S22.CrossRefGoogle ScholarPubMed
Abrahams, M. V. and Healey, M. C. (1993). Some consequences of variation in vessel density: a manipulative field experiment. Fisheries Research, 15, 315–322.CrossRefGoogle Scholar
Abramowitz, M. and Stegun, I. (1974). Handbook of Mathematical Functions. New York: Dover Publications.Google Scholar
Acosta, C. A. (2002). Spatially explicit dispersal dynamics and equilibrium population sizes in marine harvest refuges. ICES Journal of Marine Science, 59, 458–468.CrossRefGoogle Scholar
Ade, R. (1989). The Trout and Salmon Handbook. New York: Facts on File.Google Scholar
Adkison, M. D. and Zhenming, S. (2001). A comparison of salmon escapement estimates using a hierarchical Bayesian approach versus separate maximum likelihood estimation of each year's return. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1663–1671.CrossRefGoogle Scholar
Adler, F. R. (1993). Migration alone can produce persistence of host–parasitoid models. American Naturalist, 141, 642–650.CrossRefGoogle Scholar
Ahmed, A. M., Baggot, S. L., Maingon, R. and Hurd, H. (2002). The costs of mounting an immune response are reflected in the reproductive fitness of the mosquito Anopheles gambiae. Oikos, 97, 371–377.CrossRefGoogle Scholar
Allen, J. E. and Maizels, R. M. (1997). Th1-Th2: reliable paradigm or dangerous dogma?Immunology Today, 18, 387–392.CrossRefGoogle ScholarPubMed
Alpatov, W. W. and Pearl, R. (1929). Experimental studies on the duration of life. XII. Influence of temperature during the larval period and adult life on the duration of the life of the imago of drosophila melanogaster. American Naturalist, 63, 37–67.CrossRefGoogle Scholar
Anand, P. (2002). Decision-making when science is ambiguous. Science, 295, 1839.CrossRefGoogle ScholarPubMed
Anderson, J. J. (1992). A vitality-based stochastic model for organism survival. In DeAngelis, D. L. and Gross, L. J. (editors) Individual-Based Models and Approaches in Ecology. New York: Chapman and Hall, pp. 256–277.CrossRefGoogle Scholar
Anderson, J. J. (2000). A vitality-based model relating stressors and environmental properties to organismal survival. Ecological Monographs, 70, 445–470.CrossRefGoogle Scholar
Anderson, L. G. (1991a). Efficient policies to maintain total allowable catches in ITQ fisheries with at-sea processing. Land Economics, 67, 141–157.CrossRefGoogle Scholar
Anderson, L. G. (1991b). A note on market power in ITQ fisheries. Journal of Environmental Economics and Management, 21, 291–296.CrossRefGoogle Scholar
Anderson, R. M. (editor) (1982). Population Dynamics of Disease: Theory and Applications. London: Chapman and Hall.CrossRefGoogle Scholar
Anderson, R. M. (1991). The Kermack–McKendrick epidemic threshold theorem. Bulletin of Mathematical Biology, 53, 3–32.CrossRefGoogle ScholarPubMed
Anderson, R. M. and May, R. M. (1978). Regulation and stability of host–parasite population interactions. I Regulatory processes. Journal of Animal Ecology, 47, 219–247.CrossRefGoogle Scholar
Anderson, R. M. and May, R. M. (1991). Infectious Diseases of Humans. Dynamics and Control. Oxford: Oxford University Press.Google Scholar
Anonymous, (2002). The State of the World Fisheries and Aquaculture. Rome: Food and Agriculture Organization of the United Nations.
Anscombe, F. J. (1950). Sampling theory of the negative binomial and logarithmic series distributions. Biometrika, 37, 358–382.CrossRefGoogle ScholarPubMed
Antia, R. and Lipsitch, M. (1997). Mathematical models of parasite responses to host immune defenses. Parasitology, 115, S115–S167.CrossRefGoogle Scholar
Antia, R., Levin, B. R. and May, R. M. (1994). Within-host population dynamics and the evolution and maintenance of microparasite virulence. American Naturalist, 144, 457–472.CrossRefGoogle Scholar
Antia, R., Nowak, M. A. and Anderson, R. M. (1996). Antigenic variation and the within-host dynamics of parasites. Proceedings of the National Academy of Sciences, 93, 985–989.CrossRefGoogle ScholarPubMed
Apollonio, S. (1994). The use of ecosystem characteristics in fisheries management. Reviews in Fisheries Science, 2, 157–180.CrossRefGoogle Scholar
Apostolaki, P., Milner-Gulland, E. J., McAllister, M. K. and Kirkwood, G. P. (2002). Modelling the effects of establishing a marine reserve for mobile species. Canadian Journal of Fisheries and Aquatic Sciences, 59, 405–415.CrossRefGoogle Scholar
Arnold, L. (1973). Stochastic Differential Equations. New York: John Wiley and Sons.Google Scholar
Arnold, L. (1974). Stochastic Differential Equations: Theory and Applications. London: Wiley Interscience.Google Scholar
Austin, D. J., White, N. J. and Anderson, R. M. (1998). The dynamics of drug action on the within-host population growth of infectious agents: melding pharmacokinetics with pathogen population dynamics. Journal of Theoretical Biology, 194, 313–339.CrossRefGoogle ScholarPubMed
Babcock, E. A. and Pikitch, E. K. (2000). A dynamic programming model of fishing strategy choice in a multispecies trawl fishery with trip limits. Canadian Journal of Fisheries and Aquatic Sciences, 57, 357–370.CrossRefGoogle Scholar
Bailey, N. T. J. (1953). The total size of a general stochastic epidemic. Biometrika, 40, 177–185.CrossRefGoogle Scholar
Bakun, A. (1996). Patterns in the Ocean. Oceanic Process and Marine Population Dynamics. La Jolla, CA: California Sea Grant.Google Scholar
Barlow, N. D. (1993). A model for the spread of bovine Tb in New Zealand possum populations. Journal of Applied Ecology, 30, 156–164.CrossRefGoogle Scholar
Barlow, N. D. (2000). Non-linear transmission and simple models for bovine tuberculosis. Journal of Animal Ecology, 69, 703–713.CrossRefGoogle Scholar
Barnard, G. A. and Bayes, T. (1958). Studies in the history of probability and statistics. IX. Thomas Bayes's essay towards solving a problem in the doctrine of chances. Biometrika, 45, 293–315.CrossRefGoogle Scholar
Barrowman, N. J. and Myers, R. A. (2000). Still more spawner-recruitment curves: the hockey stick and its generalizations. Canadian Journal of Fisheries and Aquatic Sciences, 57, 665–676.CrossRefGoogle Scholar
Bartlett, B. R. (1964). Patterns in the host-feeding habit of adult parasite hymenoptera. Annals of the Entomological Society of America, 57, 344–350.CrossRefGoogle Scholar
Bartlett, M. S. (1957). Measles periodicity and community size. Journal of the Royal Statistical Society A, 120, 48–70.CrossRefGoogle Scholar
Basar, T. and Olsder, G. J. (1982). Dynamic Noncooperative Game Theory. New York: Academic Press.Google Scholar
Bazykin, A. D. (1998). Nonlinear Dynamics of Interacting Populations. Singapore and River Edge, NJ: World Scientific.Google Scholar
Beamish, R. J. (2002). Obituary. William Edwin Ricker OC, FRSC, LLD, DSc. Journal of Fish Biology, 60, 285–287.CrossRefGoogle Scholar
Begon, M., Bowers, R. G., Kadianakis, N. and Hodgkinson, D. E. (1992). Disease and community structure: the importance of host self-regulation in a host–host pathogen model. The American Naturalist, 139, 1131–1150.CrossRefGoogle Scholar
Begon, M., Feore, S. M., Brown, K., et al. (1998). Population and transmission dynamics of cowpox in bank voles: testing fundamental assumptions. Ecology Letters, 1, 82–86.CrossRefGoogle Scholar
Begon, M., Hazel, S. M., Baxby, D., et al. (1999). Transmission dynamics of a zoonotic pathogen within and between wildlife host species. Proceedings of the Royal Society of London, B266, 1939–1945.CrossRefGoogle ScholarPubMed
Behnke, R. J. (1992). Native Trout of Western North America. Bethesda, MD: American Fisheries Society.Google Scholar
Behnke, R. J. (2002). Trout and Salmon of North America. New York: Free Press.Google Scholar
Belew, R. K. and Mitchell, M. (editors) (1996). Adaptive Individual in Evolving Populations: Models and Algorithms. Reading, MA: Addison Wesley.Google Scholar
Bellman, R. E. (1984). Eye of the Hurricane: An Autobiography. Singapore: World Scientific.CrossRefGoogle Scholar
Bellman, R. and Cooke, K. L. (1963). Differential-Difference Equations. New York: Academic Press.Google Scholar
Bellows, T. S. and Hassell, M. P. (1988). The dynamics of age-structured host-parasitoid interactions. Journal of Animal Ecology, 57, 259–268.CrossRefGoogle Scholar
Bender, C. M. and Orszag, S. A. (1978). Advanced Mathematical Methods for Scientists and Engineers. New York: McGraw Hill.Google Scholar
Berger, R., Quack, M. and Tschumper, G. S. (2000). Electroweak quantum chemistry for possible precursor molecules in the evolution of biomolecular homochirality. Helvetica Chimica Acta, 83, 1919–1950.3.0.CO;2-D>CrossRefGoogle Scholar
Bernoulli, D. (1738/1954). Exposition of a new theory on the measurement of risk. Economietrica, 22, 23–36.CrossRefGoogle Scholar
Bertsekas, D. P (1976). Dynamic Programming and Stochastic Control. New York: Academic Press.Google Scholar
Bertsekas, D. P. (1995). Dynamic Programming and Optimal Control. Volumes One and Two. Belmont, MA: Athena Scientific.Google Scholar
Beverton, R. J. H. (1992). Patterns of reproductive strategy parameters in some marine teleost fishes. Journal of Fish Biology, 41 (Supplement B), 137–160.CrossRefGoogle Scholar
Beverton, R. J. H. (1994). Notes on the Use of Theoretical Models in the Study of the Dynamics of Exploited Fish Populations. From the Lectures of R. J. H. Beverton presented at U.S. Fishery Laboratory, Beaufort, North Carolina, Bureau of Commercial Fisheries 1951. Beaufort, NC: Marine Fisheries Section, American Fisheries Society.Google Scholar
Beverton, R. J. H. and Holt, S. J. (1957). On the Dynamics of Exploited Fish Populations. London: HMSO.Google Scholar
Beverton, R. J. H. and Holt, S. J. (1959). A review of the lifespans and mortality rates of fish in nature, and their relation to growth and other physiological characteristics. In Wolstenholme, G. E. W. and O'Connor, M. (editors) CIBA Foundation Colloquia on Ageing. London: CIBA Foundation, pp. 142–174.Google Scholar
Bharucha-Reid, A. T. (1997 (1960)). Elements of Stochastic Processes. New York: Dover.Google Scholar
Bjorksted, E. P. (2000). Stock–recruitment relationships for life cycles that exhibit concurrent density dependence. Canadian Journal of Fisheries and Aquatic Sciences, 57, 459–467.CrossRefGoogle Scholar
Bjørnstad, O. N. and Grenfell, B. T. (2001). Noisy clockwork: time series analysis of population fluctuations in animals. Science, 293, 638–643.CrossRefGoogle ScholarPubMed
Black, F. and Scholes, M. (1972). The pricing of options and corporate liabilities. The Journal of Political Economy, 81, 637–654.CrossRefGoogle Scholar
Blackmond, D. G., McMillan, C. R., Ramdeechul, S., Schorm, A. and Brown, J. M. (2001). Origins of asymmetric amplifications in autocatalytic alkylzinc additions. Journal of the American Chemical Society, 123, 10 103–10 104.CrossRefGoogle Scholar
Bleistein, N. and Handelsman, R. A. (1975). Asymptotic Expansions of Integrals. New York: Holt, Rinehart and Winston.Google Scholar
Blythe, S. P., Nisbet, R. M. and Gurney, W. S. C. (1984). The dynamics of population models with distributed maturation periods. Theoretical Population Biology, 25, 289–311.CrossRefGoogle ScholarPubMed
Bonsall, M. P. and Hastings, A. (2004). Demographic and environmental stochasticity in predator–prey metapopulation dynamics. Journal of Animal Ecology, 73, 1043–1055.CrossRefGoogle Scholar
Bonsall, M. B., Hassell, M. P. and Asefa, G. (2002). Ecological trade-offs, resource partitioning, and coexistence in a host–parasitoid assemblage. Ecology, 83, 925–934.Google Scholar
Boreman, J., Nakashima, B. S., Wilson, J. A. and Kendall, R. L. (editors) (1997). Northwest Atlantic Groundfish: Perspectives on a Fishery Collapse. Bethesda, MD: American Fisheries Society.Google Scholar
Botsford, L. W., Hastings, A. and Gaines, S. D. (2001). Dependence of sustainability on the configuration of marine reserves and larval dispersal distance. Ecology Letters, 4, 144–150.CrossRefGoogle Scholar
Bouloux, C., Langlais, M. and Silan, P. (1998). A marine host-parasite model with direct biological cycle and age structure. Ecological Modelling, 107, 73–86.CrossRefGoogle Scholar
Bradford, M., Myers, R. A. and Irvine, J. R. (2000). Reference points for coho salmon (Oncorhynchus kisutch) harvest rates and escapement goals based on freshwater production. Canadian Journal of Fisheries and Aquatic Sciences, 57, 677–686.CrossRefGoogle Scholar
Brenner, S. (1999). Theoretical biology in the third millennium. Philosophical Transactions of the Royal Society of London, Series B, Biological Sciences, 354, 1963–1965.CrossRefGoogle ScholarPubMed
Briggs, C. J. (1993). Competition among parasitoid species on a stage-structure host and its effect on host suppression. American Naturalist, 141, 372–397.CrossRefGoogle Scholar
Briggs, C. J. and Hoopes, M. F. (2004). Stabilizing effects in spatial parasitoid-host and predator-prey models: a review. Theoretical Population Biology, 65, 299–315.CrossRefGoogle ScholarPubMed
Brinkman, H. C. (1956). Brownian motion in a field of force and the diffusion theory of chemical reactions, II. Physica, 22, 149–155.CrossRefGoogle Scholar
Brodziak, J. K. T., Overholtz, W. J. and Rago, P. J. (2001). Does spawning stock affect recruitment of New England groundfish. Canadian Journal of Fisheries and Aquatic Sciences, 58, 306–318.CrossRefGoogle Scholar
Brookhart, M. A., Hubbard, A. E., Laan, M. J., Colford, J. M. Jr. and Eisenberg, J. N. S. (2002). Statistical estimation of parameters in a disease transmission model: analysis of a Cryptosporidium outbreak. Statistics in Medicine, 21, 3627–3638.CrossRefGoogle Scholar
Brooks, E. N. (2002). Using reproductive values to define optimal harvesting for multisite density-dependent populations: example with a marine reserve. Canadian Journal of Fisheries and Aquatic Sciences, 59, 875–885.CrossRefGoogle Scholar
Brown, R. (1828). A brief account of microscopical observations made in the months of June, July, and August, 1827, on the particles contained in the pollen of plants; and on the general existence of active molecules in organic and inorganic bodies. The Philosophical Magazine, 4, 161–173.CrossRefGoogle Scholar
Brunet, L. R., Dunne, D. W. and Pearce, E. J. (1998). Cytokine interaction and immune response during Schistosoma mansoni infection. Parasitology Today, 14, 422–427.CrossRefGoogle ScholarPubMed
Bucklew, J. A. (1990). Large Deviation Techniques in Decision, Simulation, and Estimation. New York: Wiley Interscience.Google Scholar
Burnham, K. P. and Anderson, D. R. (1998). Model Selection and Inference. A Practical Information–Theoretic Approach. New York: Springer Verlag.CrossRefGoogle Scholar
Bynum, W. F. (2002). Mosquitoes bite more than once. Science, 295, 47.CrossRefGoogle ScholarPubMed
Caddy, J. F. (2002). Limit reference points, traffic lights, and holistic approaches to fisheries management with minimal stock assessment input. Fisheries Research, 56, 133–137.CrossRefGoogle Scholar
Calder, W. A. I. (1984). Size, Function, and Life History. Cambridge, MA: Harvard University Press.Google Scholar
Callaway, D. S. and Perelson, A. S. (2002). HIV-1 infection and low steady state viral loads. Bulletin of Mathematical Biology, 64, 29–64.CrossRefGoogle ScholarPubMed
Caraco, T., Duryea, M. C., Glavanakov, S., Maniatty, W. and Szymanski, B. K. (2001). Host spatial heterogeneity and the spread of vector-borne infection. Theoretical Population Biology, 59, 185–206.CrossRefGoogle ScholarPubMed
Cardinale, M. and Arrhenius, F. (2000). Decreasing weight-at-age of Atlantic herring (Clupea harengus) from the Baltic Sea between 1986 and 1996: a statistical analysis. ICES Journal of Marine Science, 57, 882–893.CrossRefGoogle Scholar
Cardinale, M. and Modin, J. (1999). Changes in size-at-maturity of Baltic cod (Gadus morhua) during a period of large variations in stock size and environmental conditions. Fisheries Research, 41, 285–295.CrossRefGoogle Scholar
Carey, J. R. (2001). Insect biodemography. Annual Review of Entomology, 46, 79–110.CrossRefGoogle ScholarPubMed
Carey, J. R. (2003). Longevity. The Biology and Demography of Life Span. Princeton, NJ: Princeton University Press.Google Scholar
Carey, J. R. and Judge, D. S. (2001). Principles of biodemography with special reference to human longevity. Population: An English Selection, 13, 9–40.Google Scholar
Carpenter, S. R., Ludwig, D. and Brock, W. A. (1999). Management of eutrophication for lakes subject to potentially irreversible change. Ecological Applications, 9, 751–771.CrossRefGoogle Scholar
Carslaw, H. S. and Jaeger, J. C. (1959). Conduction of Heat in Solids. Oxford: Oxford University Press.Google Scholar
Catteruccia, F., Godfray, C. and Crisanti, A. (2003). Impact of genetic manipulation of the fitness of Anopheles stephensi mosquitoes. Science, 299, 1225–1227.CrossRefGoogle ScholarPubMed
Caughley, G. (1994). Directions in conservation biology. Journal of Animal Ecology, 63, 215–244.CrossRefGoogle Scholar
Chapman, R. N. (1928). The quantitative analysis of environmental factors. Ecology, 9, 111–122.CrossRefGoogle Scholar
Charles, A. T. (1992). Fishery conflicts. A unified framework. Marine Policy, 16, 379–393.CrossRefGoogle Scholar
Charles, S., Morand, S., Chasse, J. L. and Auger, P. (2002). Host patch selection induced by parasitism: basic reproduction ratio R0 and optimal virulence. Theoretical Population Biology, 62, 97–109.CrossRefGoogle Scholar
Charlesworth, B. (1990). Optimization models, quantitative genetics, and mutation. Evolution, 44, 520–538.CrossRefGoogle ScholarPubMed
Charlesworth, B. (1994). Evolution in Age-Structured Populations. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Charnov, E. (1993). Life History Invariants. New York: Oxford University Press.Google Scholar
Charnov, E. L. (1976). Optimal foraging, the marginal value theorem. Theoretical Population Biology, 9, 129–136.CrossRefGoogle ScholarPubMed
Charnov, E. L. and Skinner, S. W. (1984). Evolution of host selection and clutch size in parasitoid wasps. Florida Entomologist, 67, 5–21.CrossRefGoogle Scholar
Charnov, E. L. and Skinner, S. W. (1985). Complementary approaches to the understanding of parasitoid oviposition decisions. Environmental Entomology, 14, 383–391.CrossRefGoogle Scholar
Charnov, E. L. and Skinner, S. W. (1988). Clutch size in parasitoids: the egg production rate as a constraint. Evolutionary Ecology, 2, 167–174.CrossRefGoogle Scholar
Chattopadhyay, J. and Sarkar, R. R. (2003). Chaos to order: preliminary experiments with a population dynamics models of three trophic levels. Ecological Modelling, 163, 45–50.CrossRefGoogle Scholar
Cheever, J. (1978). The Stories of John Cheever. New York: Alfred A. Knopf.Google Scholar
Chen, D. G., Irvine, J. R. and Cass, A. J. (2002). Incorporating Allee effects in fish stock–recruitment models and applications for determining reference points. Canadian Journal of Fisheries and Aquatic Sciences, 59, 242–249.CrossRefGoogle Scholar
Chen, Y., Jiao, Y. and Chen, L. (2003). Developing robust frequentist and Bayesian fish stock assessment methods. Fish and Fisheries, 4, 105–120.CrossRefGoogle Scholar
Christensen, V., Guénette, S., Heymans, J. J., et al. (2003). Hundred-year decline of North Atlantic predatory fishes. Fish and Fisheries, 4, 1–24.CrossRefGoogle Scholar
Clark, C. W. (1973). The economics of overexploitation. Science, 181, 630–634.CrossRefGoogle ScholarPubMed
Clark, C. W. (1985). Bioeconomic Modelling and Fisheries Management. New York: Wiley Interscience.Google Scholar
Clark, C. W. (1990). Mathematical Bioeconomics, 2nd edn. New York: Wiley Interscience.Google Scholar
Clark, C. W. (2006). The Worldwide Crisis in Fisheries. New York: Cambridge University Press.Google Scholar
Clark, C. W. and Mangel, M. (1979). Aggregation and fishery dynamics: a theoretical study of schooling and the purse seine tuna fisheries. Fishery Bulletin, 77, 317–337.Google Scholar
Clark, C. W. and Mangel, M. (2000). Dynamic State Variable Models in Ecology. Methods and Applications. New York: Oxford University Press.Google Scholar
Clark, W. G. (1991). Groundfish exploitation rates based on life history parameters. Canadian Journal of Fisheries and Aquatic Sciences, 48, 734–750.CrossRefGoogle Scholar
Clark, W. G. (2002). F35% revisited ten years later. North American Journal of Fisheries Management, 22, 251–257.2.0.CO;2>CrossRefGoogle Scholar
Cochrane, K. L. (2000). Reconciling sustainability, economic efficiency and equity in fisheries: the one that got away?Fish and Fisheries, 1, 3–21.CrossRefGoogle Scholar
Codeço, C. T. (2000). Endemic and epidemic dynamics of cholera: the role of the aquatic reservoir. BMC Infectious Diseases, 1, 1–14.CrossRefGoogle Scholar
Cohen, D. (1966). Optimizing reproduction in a randomly varying environment. Journal of Theoretical Biology, 12, 119–129.CrossRefGoogle Scholar
Cohen, J. E. (1995). Unexpected dominance of high frequencies in chaotic nonlinear population models. Nature, 378, 610–612.CrossRefGoogle ScholarPubMed
Cole-King, A. (1993). Marine conservation: a new policy area. Marine Policy, May 1993, 171–185.CrossRefGoogle Scholar
Comins, H. N. and Hassell, M. P. (1979). The dynamics of optimally foraging predators and parasitoids. Journal of Animal Ecology, 48, 335–351.CrossRefGoogle Scholar
Comins, H. N., Hassell, M. P. and May, R. M. (1992). The spatial dynamics of host-parasitoid systems. Journal of Animal Ecology, 61, 735–748.CrossRefGoogle Scholar
Congdon, P. (2001). Bayesian Statistical Modelling. New York: John Wiley and Sons.Google Scholar
Conover, D. O. and Munch, S. B. (2002). Sustaining fisheries yields over evolutionary time scales. Science, 297, 94–96.CrossRefGoogle ScholarPubMed
Cook, R. M. and Hubbard, S. F. (1977). Adaptive searching strategies in insect parasites. Journal of Animal Ecology, 46, 115–125.CrossRefGoogle Scholar
Cooper, A. B., Hilborn, R. and Unsworth, J. W. (2003). An approach for population assessment in the absence of abundance indices. Ecological Applications, 13, 814–828.CrossRefGoogle Scholar
Cootner, P. H. (editor) (1964). The Random Character of Stock Market Prices. Cambridge, MA: MIT Press.Google Scholar
Corkett, C. J. (2002). Fish stock assessment as a non-falsifiable science: replacing an inductive and instrumental view with a critical rational one. Fisheries Research, 56, 117–123.CrossRefGoogle Scholar
Costantino, R. F. and Desharnais, R. A. (1991). Population Dynamics and the Tribolium Model: Genetics and Demography. New York: Springer Verlag.CrossRefGoogle Scholar
Cote, I. M., Mosqueira, I. and Reynolds, J. D. (2001). Effects of marine reserve characteristics on the protection of fish populations: a meta-analysis. Journal of Fish Biology, 59 (Supplement A), 178–179.CrossRefGoogle Scholar
Cottingham, K. L., Chiavelli, D. A. and Taylor, R. K. (2003). Environmental microbe and human pathogen: the ecology and microbiology of Vibrio cholerae. Frontiers in Ecology and the Environment, 1, 80–86.CrossRefGoogle Scholar
Council, N. R. (editor) (2001). Marine Protected Areas. Tools for Sustaining Ocean Ecosystems. Washington, DC: National Academy Press.Google Scholar
Courant, R. and Hilbert, D. (1962). Methods of Mathematical Physics. Volume II. Partial Differential Equations. New York: Wiley Interscience.Google Scholar
Crank, J. (1975). The Mathematics of Diffusion, 2nd edn. Oxford: Oxford University Press.Google Scholar
Crawley, M. J. (2002). Statistical Computing. An Introduction to Data Analysis Using S-Plus. New York: Wiley.Google Scholar
Crow, J. F. and Kimura, M. (1970). An Introduction to Population Genetics Theory. Minneapolis, MN: Burgess Publishing Company.Google Scholar
Crowder, L. B. and Murawski, S. A. (1998). Fisheries bycatch: Implications for management. Fisheries, 23, 8–17.2.0.CO;2>CrossRefGoogle Scholar
Daskalov, G. M. (2002). Overfishing drives a trophic cascade in the Black Sea. Marine Ecology – Progress Series, 225, 53–63.CrossRefGoogle Scholar
Davies, R. (1992). Reading and Writing. Salt Lake City: University of Utah Press.Google Scholar
Davis, H. T. (1962). Introduction to Nonlinear Differential and Integral Equations. New York: Dover.Google Scholar
Davis, P. J. and Gregerman, R. I. (1995). Occasional Notes: Parse Analysis II: A revised model that accounts for phi. The New England Journal of Medicine, 332, 965–966.CrossRefGoogle Scholar
Day, T. (2001). Parasite transmission models and the evolution of virulence. Evolution, 55, 2389–2400.CrossRefGoogle ScholarPubMed
Day, T. (2002a). The evolution of virulence in vector-borne and directly transmitted parasites. Theoretical Population Biology, 62, 199–213.CrossRefGoogle Scholar
Day, T. (2002b). On the evolution of virulence and the relationship between various measures of mortality. Proceedings of the Royal Society of London, B269, 1317–1323.CrossRefGoogle Scholar
Day, T. (2002c). Virulence evolution via host exploitation and toxin production in spore-producing pathogens. Ecology Letters, 5, 471–476.CrossRefGoogle Scholar
Day, T. (2003). Virulence evolution and the timing of disease life-history events. Trends in Ecology and Evolution, 18, 113–118.CrossRefGoogle Scholar
Day, T. and Proulx, S. (2004). A general theory for the evolutionary dynamics of virulence. American Naturalist, 164, E40–E63.CrossRefGoogle Scholar
Bach, P. (1943). The importance of host-feeding by adult parasites in the reduction of host populations. Journal of Economic Entomology, 36, 647–658.CrossRefGoogle Scholar
Bruijn, N. G. (1981). Asymptotic Methods in Analysis. New York: Dover Publications.Google Scholar
Grey, A. D. N. J. (2003a). An engineer's approach to the development of real anti-aging medicine. Science SAGE KE 2003 (8 January 2003).Google Scholar
Grey, A. D. N. J. (2003b). Overzealous maximum-likelihood fitting falsely convicts the slope heterogeneity hypothesis. Experimental Gerontology, 38, 921–923.CrossRefGoogle Scholar
Koeijer, A., Diekmann, O. and Reijnders, P. (1998). Modelling the spread of phocine distemper virus among harbour seals. Bulletin of Mathematical Biology, 60, 585–596.CrossRefGoogle ScholarPubMed
Koeijer, A., Heesterbeek, H., Schreuder, B., et al. (2004). Quantifying BSE control by calculating the basic reproduction ratio R0 for the infection among cattle. Journal of Mathematical Biology, 48, 1–22.CrossRefGoogle ScholarPubMed
Roode, J. C. and Read, A. F. (2003). Evolution and ecology, after the malaria genomes. Trends in Ecology and Evolution, 18, 60–61.CrossRefGoogle Scholar
Valpine, P. and Hastings, A. (2002). Fitting population models incorporating process noise and observation error. Ecological Monographs, 72, 57–76.CrossRefGoogle Scholar
DeAngelis, D. L. and Gross, L. J. (1992). Individual-Based Models and Approaches in Ecology. New York: Chapman and Hall.CrossRefGoogle Scholar
DeGroot, M. H. (1970). Optimal Statistical Decisions. New York: McGraw-Hill.Google Scholar
Delbrück, M. (1940). Statistical fluctuations in autocatalytic reactions. Journal of Chemical Physics, 8, 120–124.CrossRefGoogle Scholar
Demetrius, L. (2001). Mortality plateaus and directionality theory. Proceedings of the Royal Society of London, B268, 2029–2037.CrossRefGoogle ScholarPubMed
Denney, N. H., Jennings, S. and Reynolds, J. D. (2002). Life-history correlates of maximum population growth rates in marine fishes. Proceedings of the Royal Society of London, B269, 2229–2237.CrossRefGoogle ScholarPubMed
Dennis, B., Munholland, P. L. and Scott, J. M. (1991). Estimation of growth and extinction parameters for endangered species. Ecological Monographs, 61, 115–143.CrossRefGoogle Scholar
Dennis, B. and Otten, M. R. M. (2000). The joint effects of density dependence and rainfall on abundance of San Joaquin kit fox. Journal of Wildlife Management, 64, 388–400.CrossRefGoogle Scholar
Dick, E. J. (2004). Beyond ‘lognormal versus gamma’: discrimination among error distributions for generalized linear models. Fisheries Research, 70, 351–366.CrossRefGoogle Scholar
Dieckmann, U., Metz, J. A. J., Sabelis, M. W. and Sigmund, K. (2002). Adaptive dynamics of infectious diseases. In Pursuit of Virulence Management. Cambridge: Cambridge University Press.Google Scholar
Diserud, O. H. and Engen, S. (2000). A general and dynamic species abundance model, embracing the lognormal and gamma models. The American Naturalist, 155, 497–511.CrossRefGoogle ScholarPubMed
Dixit, A. K. and Pindyck, R. S. (1994). Investment Under Uncertainty. Princeton, NJ: Princeton University Press.Google Scholar
Dobson, A. and Foutopoulos, J. (2001). Emerging infectious pathogens of wildlife. Philosophical Transactions of the Royal Society London, B356, 1001–1012.CrossRefGoogle ScholarPubMed
Dobson, A. P. and Hudson, P. J. (1992). Regulation and stability of a free-living host–parasite system: Trichostrongylus tenius in red grouse. II. Population models. Journal of Animal Ecology, 61, 487–498.CrossRefGoogle Scholar
Dorn, M. W. (2002). Advice on west coast rockfish harvest rates from Bayesian meta-analysis of stock–recruit relationships. North American Journal of Fisheries Management, 22, 280–300.2.0.CO;2>CrossRefGoogle Scholar
Dovers, S. R. and Handmer, J. W. (1995). Ignorance, the precautionary principle, and sustainability. Ambio, 24, 92–97.Google Scholar
Dovers, S. R., Norton, T. W. and Handmer, J. W. (1996). Uncertainty, ecology, sustainability and policy. Biodiversity and Conservation, 5, 1143–1167.CrossRefGoogle Scholar
Driessen, G. and Hemerik, L. (1992). The time and egg budget of Leptopilina clavipes, a parasitoid of Drosophila. Ecological Entomology, 17, 17–27.CrossRefGoogle Scholar
Dubins, L. E. and Savage, L. J. (1976). Inequalities for Stochastic Processes: How to Gamble if You Must. New York: Dover Publications.Google Scholar
Dwyer, G. and Elkinton, J. S. (1993). Using simple models to predict virus epizootics in gypsy moth populations. Journal of Animal Ecology, 62, 1–11.CrossRefGoogle Scholar
Dyson, F. (1999). The Sun, the Genome, and the Internet. New York: Oxford University Press.Google Scholar
Earn, D. J. D., Dushoff, J. and Levin, S. A. (2002). Ecology and evolution of the flu. Trends in Ecology and Evolution, 17, 334–340.CrossRefGoogle Scholar
Easterling, M. R. and Ellner, S. P. (2000). Dormancy strategies in a random environment: comparing structured and unstructured models. Evolutionary Ecology Research, 2, 387–407.Google Scholar
Ebert, D. and Bull, J. J. (2003). Challenging the trade-off model for the evolution of virulence: is virulence management feasible?Trends in Microbiology, 11, 15–20.CrossRefGoogle ScholarPubMed
Edelstein-Keshet, L. (1988). Mathematical Models in Biology. New York: Random House.Google Scholar
Edgar, G. and Barrett, N. S. (1999). Effects of the declaration of marine reserves on Tasmanian reef fishes, invertebrates and plants. Journal of Experimental Marine Biology and Ecology, 242, 107–144.CrossRefGoogle Scholar
Edwards, A. W. F. (1992). Likelihood, expanded edition. Baltimore: Johns Hopkins University Press.Google Scholar
Edwards, R. L. (1954). The effect of diet on egg maturation and resorption in Mormoniella vitripennis (Hymenoptera, Pteromalidae). Quarterly Journal of Microscopical Science, 95, 459–469.Google Scholar
Efron, B. (2005). Bayesians, frequentists, and scientists. Journal of the American Statistical Association, 100, 1–5.CrossRefGoogle Scholar
Eigen, M. (1996). Prionics or the kinetic basis of prion diseases. Biophysical Chemistry, 63, A1–A18.CrossRefGoogle ScholarPubMed
Einstein, A. (1956). Investigations on the Theory of Brownian Movement. New York: Dover Publications.Google Scholar
Einum, S. and Fleming, I. A. (2000). Highly fecund mothers sacrifice offspring survival to maximize fitness. Nature, 405, 565–567.CrossRefGoogle ScholarPubMed
Ellison, A. M. (1996). An introduction to Bayesian inference for ecological research and environmental decision-making. Ecological Applications, 6, 1036–1046.CrossRefGoogle Scholar
Ellison, A. M. (2004). Bayesian inference in ecology. Ecology Letters, 7, 509–520.CrossRefGoogle Scholar
Emlen, J. M. (1966). The role of time and energy in food preference. American Naturalist, 100, 611–617.CrossRefGoogle Scholar
Enderby, J. E. (1998). Sir Frederick Charles Frank (1911–98). Nature, 393, 314.CrossRefGoogle Scholar
Engen, S., Lande, R. and Saether, B.-E. (2002). The spatial scale of population fluctuations and quasi-extinction risk. American Naturalist, 160, 439–451.Google ScholarPubMed
Engen, S. and Saether, B.-E. (2000). Predicting the time of quasi-extinction for populations far below their carrying capacity. Journal of Theoretical Biology, 205, 649–658.CrossRefGoogle ScholarPubMed
Engen, S., Saether, B.-E. and Møller, A. P. (2001). Stochastic population dynamics and time to extinction of a declining population of barn swallows. Journal of Animal Ecology, 70, 789–797.CrossRefGoogle Scholar
Enquist, B. J., West, G. B., Charnov, E. L. and Brown, J. H. (1999). Allometric scaling of production and life-history variation in vascular plants. Nature, 401, 907–911.CrossRefGoogle Scholar
Enserink, M. (2002). How devastating would a smallpox attack really be?Science, 296, 1592–1595.CrossRefGoogle Scholar
Enserink, M. (2003). New look at old data irks smallpox-eradication experts. Science, 299, 181.CrossRefGoogle Scholar
Esch, G. W., Shostak, A. W., Marcogliese, D. J. and Goater, T. M. (1990). Parasite Communities: Patterns and Processes. New York: Chapman and Hall.Google Scholar
Essington, T. E., Kitchell, J. F. and Walters, C. J. (2001). The von Bertalanffy growth function, bioenergetics, and the consumption rates of fish. Canadian Journal of Fisheries and Aquatic Sciences, 58, 2129–2138.CrossRefGoogle Scholar
Estes, J. A., Tinker, M. T., Williams, T. M. and Doak, D. F. (1998). Killer whale predation on sea otters linking oceanic and nearshore ecosystems. Science, 282, 473–476.CrossRefGoogle ScholarPubMed
Esteva, L. and Vargas, C. (2003). Coexistence of different serotypes of dengue virus. Journal of Mathematical Biology, 46, 31–47.CrossRefGoogle ScholarPubMed
Ewald, P. W. (1994). Evolution of Infectious Disease. New York: Oxford University Press.Google Scholar
Fanshawe, S., Vanblaricom, G. R. and Shelly, A. A. (2003). Restored top carnivores as detriments to the performance of marine protected areas intended for fishery sustainability: a case study with red abalones and sea otters. Conservation Biology, 17, 273–283.CrossRefGoogle Scholar
Farrow, S. and Sumaila, U. R. (2002). Economics of marine protected areas. Fish and Fisheries, 3, 356–359.Google Scholar
Faruque, S., Chowdhury, N., Kamruzzaman, M.et al. (2003). Reemergence of Epidemic Vibrio cholerae O139, Bangladesh. Emerging Infectious Diseases, 9, 1116–1122.CrossRefGoogle ScholarPubMed
Feller, W. (1957). An Introduction to Probability Theory and Its Applications, Volume I. New York: John Wiley and Sons.Google Scholar
Feller, W. (1971). An Introduction to Probability Theory and Its Applications, Volume II. New York: John Wiley and Sons.Google Scholar
Ferguson, H. M. and Read, A. F. (2002). Genetic and environmental determinants of malaria parasite virulence in mosquitoes. Proceedings of the Royal Society of London, B269, 1217–1224.CrossRefGoogle ScholarPubMed
Ferguson, N. M., Ghani, A. C., Donnelly, C. A., Hagenaars, T. J. and Anderson, R. M. (2002). Estimating the human health risk from possible BSE infection of the British sheep flock. Nature, 415, 420–424.CrossRefGoogle ScholarPubMed
Feynman, R. P. (1948). Space-time approach to non-relativistic quantum mechanics. Reviews of Modern Physics, 20, 367–387.CrossRefGoogle Scholar
Feynman, R. P. (1985). Surely You're Joking Mr. Feynman. New York: W. W. Norton.Google Scholar
Fieberg, J. and Ellner, S. P. (2000). When is it meaningful to estimate an extinction probability?Ecology, 81, 2040–2047.CrossRefGoogle Scholar
Fiksen, Ø. and Slotte, A. (2002). Stock-environment recruitment models for Norwegian spring spawning herring (Clupea harengus). Canadian Journal of Fisheries and Aquatic Sciences, 59, 211–217.CrossRefGoogle Scholar
Fisher, A. C., Hanemann, W. M. and Keeler, A. G. (1991). Integrating fishery and water resource management: a biological model of a California salmon fishery. Journal of Environmental Economics and Management, 20, 234–261.CrossRefGoogle Scholar
Fisher, R. A. (1930) (1958 reprint). The Genetical Theory of Natural Selection, 2nd revised edn. New York: Dover.CrossRefGoogle Scholar
Fishman, M. A. and Perelson, A. (1999). Th1/Th2 differentiation and cross-regulation. Bulletin of Mathematical Biology, 61, 403–436.CrossRefGoogle ScholarPubMed
Flanders, S. E. (1950). Regulation of ovulation and egg disposal in the parasitic hymenoptera. The Canadian Entomologist, 82, 134–140.CrossRefGoogle Scholar
Flather, C. H. (1996). Fitting species-accumulation functions and assessing regional land use impacts on avian diversity. Journal of Biogeography, 23, 155–168.CrossRefGoogle Scholar
Fletcher, J. P., Hughes, J. P. and Harvey, I. F. (1994). Life expectancy and egg load affect oviposition decision of a solitary parasitoid. Proceedings of the Royal Society of London, B258, 163–167.CrossRefGoogle Scholar
Foley, P. (1994). Predicting extinction times from environmental stochasticity and carrying capacity. Conservation Biology, 8, 124–137.CrossRefGoogle Scholar
Francis, R. I. C. C. and Shotton, R. (1997). “Risk” in fisheries management: a review. Canadian Journal of Fisheries and Aquatic Sciences, 54, 1699–1715.Google Scholar
Frank, F. C. (1953). On spontaneous asymmetric synthesis. Biochimica et Biophysica Acta, 11, 459–463.CrossRefGoogle ScholarPubMed
Frank, S. A. (1996). Models of parasite virulence. The Quarterly Review of Biology, 71, 37–78.CrossRefGoogle ScholarPubMed
Frank, S. A. (1998). Social Evolution. Princeton, NJ: Princeton University Press.Google ScholarPubMed
Frank, S. A. (2002). Immunology and Evolution of Infectious Diseases. Princeton, NJ: Princeton University Press.Google Scholar
Freedman, D., Pisani, R. and Purves, R. (1998). Statistics, 3rd edn. New York: W. W. Norton.Google Scholar
Friedlin, M. I. and Wentzell, A. D. (1984). Random Perturbations of Dynamical Systems. New York: Springer Verlag.CrossRefGoogle Scholar
Frisk, M. G., Miller, T. J. and Fogarty, M. J. (2001). Estimation and analysis of biological parameters in elasmobranch fishes: a comparative life history study. Canadian Journal of Fisheries and Aquatic Sciences, 58, 969–981.CrossRefGoogle Scholar
Fuerth, R. (1956). Notes. In Investigations on the Theory of the Brownian Movement by Albert Einstein, Ph.D. New York: Dover Publications, pp. 86–119.Google Scholar
Fulford, G. R., Roberts, M. G. and Heesterbeek, J. A. P. (2002). The metapopulation dynamics of an infectious disease: tuberculosis in possums. Theoretical Population Biology, 61, 15–29.CrossRefGoogle ScholarPubMed
Galvani, A. P. (2003). Epidemiology meets evolutionary ecology. Trends in Ecology and Evolution, 18, 132–139.CrossRefGoogle Scholar
Gandon, S., Mackinnon, M. J., Nee, S. and Read, A. F. (2001). Imperfect vaccines and the evolution of pathogen virulence. Nature, 414, 751–755.CrossRefGoogle ScholarPubMed
Gani, R. and Leach, S. (2001). Transmission potential of smallpox in contemporary populations. Nature, 414, 748–751.CrossRefGoogle ScholarPubMed
Ganusov, V. V., Bergstrom, C. T. and Antia, R. (2002). Within-host population dynamics and the evolution of microparasites in a heterogeneous host population. Evolution, 56, 213–223.CrossRefGoogle Scholar
Gardiner, C. W. (1983). Handbook of Stochastic Methods for Physics, Chemistry and the Natural Sciences. Berlin: Springer Verlag.CrossRefGoogle Scholar
Gardner, S. N. (2000). Scheduling chemotherapy: Catch 22 between cell kill and resistance evolution. Journal of Theoretical Medicine, 2, 21–232.CrossRefGoogle Scholar
Gardner, S. N. (2001). Modeling multi-drug chemotherapy: tailoring treatment to individuals. Journal of Theoretical Biology, 214, 181–207.CrossRefGoogle Scholar
Gardner, S. N. and Agrawal, A. A. (2002). Induced plant defense and the evolution of counter-defenses in herbivores. Evolutionary Ecology Research, 4, 1131–1151.Google Scholar
Gardner, S. N. and Thomas, M. B. (2002). Costs and benefits of fighting infection in locusts. Evolutionary Ecology Research, 4, 109–131.Google Scholar
Gavrilets, S. (2003). Models of speciation: what have we learned in 40 years?Evolution, 57, 2197–2215.CrossRefGoogle ScholarPubMed
Gavrilov, L. A. and Gavrilova, N. S. (1991). The Biology of Life Span: A Quantitative Approach. London: Harwood Academic Publishers.Google Scholar
Gelman, A., Carlin, J. B., Stern, H. S. and Rubin, D. B. (1995). Bayesian Data Analysis. London: Chapman and Hall.Google Scholar
George, E. (2004). Write Away. New York: Harper Collins.Google Scholar
Gillespie, D. T. (1992). Markov Processes. An Introduction for Physical Scientists. Boston: Academic Press.Google Scholar
Gillespie, D. T. (1996). Exact numerical simulation of the Ornstein–Uhlenbeck process and its integral. Physical Review E, 54, 2084–2091.CrossRefGoogle ScholarPubMed
Gillespie, J. H. (1991). The Causes of Molecular Evolution. New York: Oxford University Press.Google Scholar
Gillespie, J. H. (1998). Population Genetics. A Concise Guide. Baltimore: Johns Hopkins University Press.Google Scholar
Gillig, D., Griffin, W. L. and Ozuna, T. J. (2001). A bioeconomic assessment of Gulf of Mexico red snapper management policies. Transactions of American Fisheries Society, 130, 117–129.2.0.CO;2>CrossRefGoogle Scholar
Gillis, D. M. (1999). Behavioral inferences from regulatory observer data: catch rate variation in the Scotian Shelf silver hake (Merluccius bilinearis) fishery. Canadian Journal of Fisheries and Aquatic Sciences, 56, 288–296.CrossRefGoogle Scholar
Gillis, D. M. (2003). Ideal free distributions in fleet dynamics: a behavioral perspective on vessel movement in fisheries analysis. Canadian Journal of Zoology, 81, 177–187.CrossRefGoogle Scholar
Gillis, D. M. and Peterman, R. M. (1998). Implications of interference among fishing vessels and the ideal free distribution to the interpretation of CPUE. Canadian Journal of Fisheries and Aquatic Sciences, 55, 37–46.CrossRefGoogle Scholar
Gillis, D. M., Pikitch, E. K. and Peterman, R. M. (1995). Dynamic discarding decisions: foraging theory for high-grading in a trawl fishery. Behavioral Ecology, 6, 146–154.CrossRefGoogle Scholar
Giron, D., Rivero, A., Mandon, N., Darrouzet, E. and Casas, J. (2002). The physiology of host feeding in parasitic wasps: implications for survival. Functional Ecology, 16, 750–757.CrossRefGoogle Scholar
Giske, J., Mangel, M., Jakobsen, P., et al. (2002). Explicit trade-off rules in proximate adaptive agents. Evolutionary Ecology Reseach, 5, 835–865.Google Scholar
Gislason, H. (1994). Ecosystem effects of fishing activities in the North Sea. Marine Pollution Bulletin, 29, 520–527.CrossRefGoogle Scholar
Gleick, J. (1988). Chaos. London: William Heinneman Ltd.Google Scholar
Gleick, J. (1992). Genius. The Life and Science of Richard Feynman. New York: Pantheon.Google Scholar
Godfray, H. C. J. (1994). Parasitoids. Behavioral and Evolutionary Ecology. Princeton, NJ: Princeton University Press.Google Scholar
Godfray, H. C. J. and Hassell, M. P. (1989). Discrete and continuous insect populations in tropical environments. Journal of Animal Ecology, 58, 153–174.CrossRefGoogle Scholar
Goel, N. S. and Richter-Dyn, N. (1974). Stochastic Models in Biology. New York: Academic Press.Google Scholar
Gompertz, B. (1825). On the nature of the function expressive of the law of human mortality, and on a new mode of determining the value of life contingencies. Philosophical Transactions of the Royal Society, 115, 513–583.CrossRefGoogle Scholar
Gordon, D. M., Nisbet, R. M., Roos, A., Gurney, W. S. C. and Stewart, R. K. (1991). Discrete generations in host–parasitoid models with contrasting life cycles. Journal of Animal Ecology, 60, 295–308.CrossRefGoogle Scholar
Gordon, H. S. (1954). The economic theory of a common property resource: the fishery. Journal of Political Economy, 62, 124–142.CrossRefGoogle Scholar
Gotelli, N. J. (2001). A Primer of Ecology, 3rd edn. Sunderland, MA: Sinauer Associates.Google Scholar
Gould, S. J. (2002). The Structure of Evolutionary Theory. Cambridge, MA: The Belknap Press of Harvard University Press.Google Scholar
Gould, S. J. and Eldredge, N. (1977). Punctuated equilibria: the tempo and mode of evolution reconsidered. Paleobiology, 3, 115–151.CrossRefGoogle Scholar
Gould, S. J. and Eldredge, N. (1993). Punctuated equilibrium comes of age. Nature, 366, 223–227.CrossRefGoogle ScholarPubMed
Grafen, A. and Hails, R. S. (2002). Modern Statistics for the Life Sciences. Learn How to Analyse your Experiments. Oxford: Oxford University Press.Google Scholar
Graham, A. L. (2001). Use of an optimality model to solve the immunological puzzle of concomitant infection. Parasitology, 122, S61–S64.CrossRefGoogle ScholarPubMed
Graham, A. L. (2002). When T-helper cells don't help: immunopathology during concomitant infection. Quarterly Review of Biology, 77, 409–434.CrossRefGoogle ScholarPubMed
Gray, R. D. (1987). Faith and foraging: a critique of the “paradigm argument from design”. In Kamil, A. C., Krebs, J. R. and Pulliam, H. R. (editors) Foraging Behavior. New York: Plenum, pp. 69–140.CrossRefGoogle Scholar
Greenberg, R. (2001). Lecture 38: Nineteenth-Century Italian Opera – Giuseppe Verdi. In How to Listen to and Understand Great Music. The Teaching Company.Google Scholar
Greene, C. (2000). Habitat selection reduces extinction of populations subject to Allee effects. Theoretical Population Biology, 64, 1–10.CrossRefGoogle Scholar
Greenwood, M. and Yule, G. U. (1920). An inquiry into the nature of frequency distributions representative of multiple happenings with particular references to the occurrence of multiple attacks of disease or of repeated accidents. Journal of the Royal Statistical Society, 83, 255–279.CrossRefGoogle Scholar
Greer, R. (1995). Ferox Trout and Arctic Charr. A Predator, its Pursuit and its Prey. Shrewsbury, UK: Swan Hill Press.Google Scholar
Grenfell, B. T. (1988). Gastrointestinal nematode parasites and the stability and productivity of intensive ruminant grazing systems. Philosophical Transactions of the Royal Society of London, B321, 541–563.CrossRefGoogle ScholarPubMed
Grenfell, B. T. (1992). Parasitism and the dynamics of ungulate grazing systems. American Naturalist, 139, 907–929.CrossRefGoogle Scholar
Grindrod, P. (1996). The Theory and Applications of Reaction–Diffusion Equations. Patterns and Waves. Oxford: Oxford University Press.Google Scholar
Groot, C. and Margolis, L. (editors) (1991). Pacific Salmon Life Histories. Vancouver, BC: University of British Columbia Press.Google Scholar
Groot, C., Margolis, L. and Clarke, W. C. (editors) (1995). The Physiological Ecology of Pacific Salmon. Vancouver, BC: University of British Columbia Press.Google Scholar
Guedj, D. (2000). The Parrot's Theorem. New York: St. Martin's Griffin.Google Scholar
Guenette, S. and Pitcher, T. J. (1999). An age-structured model showing the benefits of marine reserves in controlling overexploitation. Fisheries Research, 39, 295–303.CrossRefGoogle Scholar
Guenette, S., Lauck, T. and Clark, C. (1998). Marine reserves: from Beverton and Holt to the present. Reviews in Fish Biology and Fisheries, 8, 251–272.CrossRefGoogle Scholar
Gulland, J. A. (editor) (1977). Fish Population Dynamics. Chichester: John Wiley and Sons.Google Scholar
Gulland, J. A. (1988). Fish Population Dynamics. Implications for Management, 2nd edn. Chichester: John Wiley & Sons.Google Scholar
Gunderson, D. R. (1997). Trade-off between reproductive effort and adult survival in oviparous and viviparous fishes. Canadian Journal of Fisheries and Aquatic Sciences, 54, 990–998.CrossRefGoogle Scholar
Gurney, W. S. C., Nisbet, R. M. and Lawton, J. H. (1983). The systematic formulation of tractable single-species population models incorporating age structure. Journal of Animal Ecology, 52, 479–495.CrossRefGoogle Scholar
Haberman, R. (1998). Elementary Applied Partial Differential Equations with Fourier Series and Boundary Value Problems. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Haccou, P., Vlas, S. J., Alhen, J. J. M. and Visser, M. E. (1991). Information processing by foragers: effects of intra-patch experience on the leaving tendency of Leptopilina heterotoma. Journal of Animal Ecology, 60, 93–106.CrossRefGoogle Scholar
Hadamard, J. (1954). The Psychology of Invention in the Mathematical Field. New York: Dover.Google Scholar
Hakoyama, H. and Iwasa, Y. (2000). Extinction risk of a density-dependent population estimated from a time series of population size. Journal of Theoretical Biology, 204, 337–359.CrossRefGoogle ScholarPubMed
Hakoyama, H., Iwasa, Y. and Nakanishi, J. (2000). Comparing risk factors for population extinction. Journal of Theoretical Biology, 204, 327–336.CrossRefGoogle ScholarPubMed
Halley, J. M. and Iwasa, Y. (1998). Extinction rate of a population under both demographic and environmental stochasticity. Theoretical Population Biology, 53, 1–15.CrossRefGoogle ScholarPubMed
Halpern, B. S. and Warner, R. R. (2002). Marine reserves have rapid and lasting effects. Ecology Letters, 5, 361–366.CrossRefGoogle Scholar
Ham, K. D. and Pearsons, T. N. (2000). Can reduced salmonid population abundance be detected in time to limit management impacts?Canadian Journal of Fisheries and Aquatic Sciences, 57, 17–24.CrossRefGoogle Scholar
Hamilton, W. D. (1966). The moulding of senescence by natural selection. Journal of Theoretical Biology, 12, 12–45.CrossRefGoogle ScholarPubMed
Hamilton, W. D. (1967). Extraordinary sex ratios. Science, 156, 477–488.CrossRefGoogle ScholarPubMed
Hamilton, W. D. (1995). Narrow Roads of Gene Land. Volume 1. Evolution of Social Behaviour. New York: W. H. Freeman and Company.Google Scholar
Hammersley, J. M. (1974). Statistical tools. The Statistician, 23, 89–106.CrossRefGoogle Scholar
Hammond, T. R. and O'Brien, C. M. (2001). An application of the Bayesian approach to stock assessment model uncertainty. ICES Journal of Marine Science, 58, 648–656.CrossRefGoogle Scholar
Hanski, I. (1989). Metapopulation dynamics: does it help to have more of the same?Trends in Ecology and Evolution, 4, 113–114.CrossRefGoogle ScholarPubMed
Hanski, I. (1999). Metapopulation Ecology. Oxford: Oxford University Press.Google Scholar
Hare, S. R. and Francis, R. C. (1995). Climate change and salmon production in the Northeast Pacific Ocean. Canadian Special Publications of Fisheries and Aquatic Sciences, 121, 357–372.Google Scholar
Harley, S. J. and Myers, R. A. (2001). Hierarchial Bayesian models of length-specific catchability of research trawl surveys. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1569–1584.CrossRefGoogle Scholar
Harley, S. J., Myers, R. A. and Dunn, A. (2001). Is catch-per-unit effort proportional to abundance?Canadian Journal of Fisheries and Aquatic Sciences, 58, 1760–1772.CrossRefGoogle Scholar
Harris, P. J. and Dean, J. M. (1998). Characterization of king mackerel and Spanish mackerel bycatches of South Carolina shrimp trawlers. North American Journal of Fisheries Management, 18, 439–453.2.0.CO;2>CrossRefGoogle Scholar
Harrison, L. G. (1993). Kinetic Theory of Living Pattern. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Hart, P. J. B. and Reynolds, J. D. (editors) (2002). Handbook of Fish Biology and Fisheries, Volume 2. Fisheries. Oxford: Blackwell.CrossRefGoogle Scholar
Harte, J. (1988). Consider a Spherical Cow. A Course in Environmental Problem Solving. Sausalito, CA: University Science Books.Google Scholar
Harte, J. (2001). Consider a Cylindrical Cow. More Adventures in Environmental Problem Solving. Sausalito, CA: University Science Books.Google Scholar
Hassell, M. P. (1978). The Dynamics of Arthropod Predator Prey Systems. Princeton, NJ: Princeton University Press.Google ScholarPubMed
Hassel, M. P. (2000a). The Spatial and Temporal Dynamics of Host–Parasitoid Interactions. Oxford: Oxford University Press.Google Scholar
Hassell, M. P. (2000b). Host–parasitoid population dynamics. Journal of Animal Ecology, 69, 543–566.CrossRefGoogle Scholar
Hassell, M. P. and May, R. M. (1988). Spatial heterogeneity and the dynamics of parasitoid–host systems. Annals Zoologia Fennici, 25, 55–61.Google Scholar
Hassell, M. P., Waage, J. K. and May, R. M. (1983). Variable parasitoid sex ratios and their effect on host-parasitoid dynamics. Journal of Animal Ecology, 52, 889–904.CrossRefGoogle Scholar
Hastings, A. and Botsford, L. W. (1999). Equivalence in yield from marine reserves and traditional fisheries management. Science, 284, 1537–1538.CrossRefGoogle ScholarPubMed
Hastings, I. M. (1997). A model for the origins and spread of drug-resistant malaria. Parasitology, 115, 133–141.CrossRefGoogle ScholarPubMed
Hastings, I. M. (2001). Modelling parasite drug resistance: lessons for management and control strategies. Tropical Medicine and International Health, 6, 883–890.CrossRefGoogle ScholarPubMed
Hastings, I. M. and D'Allesandro, U. (2000). Modelling a predictable disaster: the rise and spread of drug-resistant malaria. Parasitology Today, 16, 340–347.CrossRefGoogle Scholar
Hastings, I. M. and MacKinnon, M. J. (1998). The emergence of drug-resistant malaria. Parasitology, 117, 411–417.CrossRefGoogle ScholarPubMed
Hastings, I. M., Bray, P. G. and Ward, S. A. (2002). A requiem for chloroquine. Science, 297, 74–75.CrossRefGoogle Scholar
Haupt, R. L. and Haupt, S. E. (1998). Practical Genetic Algorithms. New York: John Wiley and Sons.Google Scholar
He, J. X. and Stewart, D. J. (2001). Age and size at first reproduction of fishes: predictive models based only on growth trajectories. Ecology, 82, 784–791.CrossRefGoogle Scholar
Healey, M. C. (1985). Influence of fishermen's preference on the success of commercial fishery management regimes. North American Journal of Fisheries Management, 5, 173–180.2.0.CO;2>CrossRefGoogle Scholar
Healey, M. C. (1990). Implications of climate change for fisheries management policy. Transactions of American Fisheries Society, 119, 366–373.2.3.CO;2>CrossRefGoogle Scholar
Healey, M. C. and Morris, J. F. T. (1992). The relationship between the dispersion of salmon fishing vessels and their catch. Fisheries Research, 15, 135–145.CrossRefGoogle Scholar
Heath, D. D., Heath, J. W., Bryden, C. A., Johnson, R. M. and Fox, C. W. (2003). Rapid evolution of egg size in captive salmon. Science, 299, 1738–1740.CrossRefGoogle ScholarPubMed
Hedrick, P. W., Lacy, R. C., Allendorf, F. W. and Soule, M. E. (1996). Directions in conservation biology. Conservation Biology, 10, 1312–1320.CrossRefGoogle Scholar
Heimpel, G. E. and Collier, T. R. (1996). The evolution of host-feeding behaviour in parasitoids. Biological Reviews, 71, 373–400.CrossRefGoogle Scholar
Heimpel, G. E. and Rosenheim, J. A. (1995). Dynamic host feeding by the parasitoid Aphytis melinus: the balance between current and future reproduction. Journal of Animal Ecology, 64, 153–167.CrossRefGoogle Scholar
Heimpel, G. E. and Rosenheim, J. A. (1998). Egg limitation in parasitoids: a review of the evidence and a case study. Biological Control, 11, 160–168.CrossRefGoogle Scholar
Heimpel, G. E., Mangel, M. and Rosenheim, J. A. (1998). Effects of time limitation and egg limitation on lifetime reproductive success of a parasitoid in the field. American Naturalist, 152, 273–289.CrossRefGoogle ScholarPubMed
Heimpel, G. E., Rosenheim, J. A. and Adams, J. A. (1994). Behavioral ecology of host feeding in Aphytis parasitoids. Norwegian Journal of Agricultural Sciences, Supplement, 16, 101–115.Google Scholar
Heimpel, G. E., Rosenheim, J. A. and Mangel, M. (1996). Egg limitation, host quality, and dynamic behavior by a parasitoid in the field. Ecology, 77, 2410–2420.CrossRefGoogle Scholar
Heimpel, G. E., Rosenheim, J. A. and Kattari, D. (1997). Adult feeding and lifetime reproductive success in the parasitoid Aphytis melinus. Entomologia Experimentalis et Applicata, 83, 305–315.CrossRefGoogle Scholar
Hellriegel, B. (2001). Immunoepidemiology – bridging the gap between immunology and epidemiology. Trends in Parasitology, 17, 102–106.CrossRefGoogle ScholarPubMed
Helser, T. E., Thunberg, E. M. and Mayo, R. K. (1996). An age-structured bioeconomic simulation of U.S. Silver hake Fisheries. North American Journal of Fisheries Management, 16, 783–794.2.3.CO;2>CrossRefGoogle Scholar
Hemerik, L., Driessen, G. and Haccou, P. (1993). Effects of intra-patch experiences on patch time, search time and searching efficiency of the parasitoid Leptopilina clavipes. Journal of Animal Ecology, 62, 33–44.CrossRefGoogle Scholar
Hemerik, L., Hoeven, N. and Alphen, J. J. M. (2002). Egg distributions and the information a solitary parasitoid has and uses for its oviposition decisions. Acta Biotheoretica, 50, 167–188.CrossRefGoogle ScholarPubMed
Henderson, D. A. (1999). The looming threat of bioterrorism. Science, 283, 1279–1282.CrossRefGoogle ScholarPubMed
Hendry, A. P. and Kinnison, M. T. (2001). An introduction to microevolution: rate, pattern, process. Genetica, 112–113, 1–8.CrossRefGoogle ScholarPubMed
Hennemuth, R. C., Palmer, J. E. and Brown, B. E. (1980). A statistical description of recruitment in eighteen selected fish stocks. Journal of Northwest Atlantic Fishery Science, 1, 101–111.CrossRefGoogle Scholar
Hernandez, M.-J. and Barradas, I. (2003). Variation in the outcome of population interactions: bifurcations and catastrophes. Journal of Mathematical Biology, 46, 571–594.CrossRefGoogle ScholarPubMed
Hershatter, G. (1997). Dangerous Pleasures. Berkeley, CA: University of California Press.Google Scholar
Hess, G. (1996). Disease in metapopulation models: implications for conservation. Ecology, 77, 1617–1632.CrossRefGoogle Scholar
Hethcote, H. W. (2000). The mathematics of infectious diseases. SIAM Review, 42, 599–653.CrossRefGoogle Scholar
Higham, D. J. (2001). An algorithmic introduction to numerical simulation of stochastic differential equations. SIAM Review, 43, 525–546.CrossRefGoogle Scholar
Highman, N. J. (1998). Handbook of Writing for the Mathematical Sciences, 2nd edn. Philadelphia, PA: SIAM (Society for Industrial and Applied Mathematics).Google Scholar
Hilborn, R. and Mangel, M. (1997). The Ecological Detective. Confronting Models with Data. Princeton, NJ: Princeton University Press.Google Scholar
Hilborn, R., Parama, A. and Maunder, M. (2002). Exploitation rate reference points for west coast rockfish: are they robust and are there better alternatives?North American Journal of Fisheries Management, 22, 365–375.2.0.CO;2>CrossRefGoogle Scholar
Hines, W. G. S. (1987). Evolutionarily stable strategies: a review of basic theory. Theoretical Population Biology, 31, 195–272.CrossRefGoogle ScholarPubMed
Hinrichsen, R. A. (2002). The accuracy of alternative stochastic growth rate estimates for salmon populations. Canadian Journal of Fisheries and Aquatic Sciences, 59, 1014–1023.CrossRefGoogle Scholar
Hochberg, M. E. and Ives, A. R. (editors) (2000). Parasitoid Population Biology. Princeton, NJ: Princeton University Press.Google Scholar
Hofbauer, J. and Sigmund, K. (1998). Evolutionary Games and Population Dynamics. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Holden, A. (editor) (2001). The New Penguin Opera Guide. London: Penguin Books.Google Scholar
Holland, D. S. and Sutinen, J. G. (1999). An empirical model of fleet dynamics in New England trawl fisheries. Canadian Journal of Fisheries and Aquatic Sciences, 56, 253–264.CrossRefGoogle Scholar
Holoman, D. K. (1992). Evenings with the Orchestra. New York: Norton.Google Scholar
Horwood, J. W., Nichols, J. H. and Milligan, S. (1998). Evaluation of closed areas for fish stock conservation. Journal of Applied Ecology, 35, 893–903.CrossRefGoogle Scholar
Hotelling, H. (1951). The impact of R. A. Fisher on statistics. Journal of the American Statistical Association, 46, 35–46.CrossRefGoogle Scholar
Houston, A. I. and McNamara, J. M. (1999). Models of Adaptive Behaviour. Cambridge: Cambridge University Press.Google Scholar
Hubbard, S. F. and Cook, R. M. (1978). Optimal foraging by parasitoid wasps. Journal of Animal Ecology, 47, 593–604.CrossRefGoogle Scholar
Hubbard, S. F., Marris, G. and Reynolds, D. M. (1987). Adaptive patterns in the avoidance of superparasitism by solitary parasitic wasps. Journal of Animal Ecology, 56, 387–401.CrossRefGoogle Scholar
Hubbell, S. (2001). The Unified Neutral Theory of Biodiversity and Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
Hudson, D. J. (1971). Interval estimation from the likelihood function. Journal of the Royal Statistical Society, Series B, 33, 256–262.Google Scholar
Hudson, P. J. (1986). The effect of a parasitic nematode on the breeding production of red grouse. Journal of Animal Ecology, 55, 85–92.CrossRefGoogle Scholar
Hudson, P. J. and Dobson, A. P. (1997). Transmission dynamics and host–parasite interactions of Trichostrongylus tenuis in red grouse (Lagopus lagopus scoticus). Journal of Parasitology, 83, 194–202.CrossRefGoogle Scholar
Hudson, P. J., Dobson, A. P. and Newborn, D. (1992a). Do parasites make prey vulnerable to predation – red grouse and parasites. Journal of Animal Ecology, 61, 681–692.CrossRefGoogle Scholar
Hudson, P. J., Newborn, D. and Dobson, A. P. (1992b). Regulation and stability of a free-living host–parasite system – Trichostrongylus tenuis in red grouse. 1. Monitoring and parasite reduction experiments. Journal of Animal Ecology, 61, 477–486.CrossRefGoogle Scholar
Hudson, P. J., Dobson, A. P. and Newborn, D. (1998). Prevention of population cycles by parasite removal. Science, 282, 2256–2258.CrossRefGoogle ScholarPubMed
Hudson, P. J., Dobson, A. P., Cattadori, I. M., et al. (2002). Trophic interactions and population growth rates: describing patterns and identifying mechanisms. Philosophical Transactions of the Royal Society of London, B357, 1259–1271.CrossRefGoogle ScholarPubMed
Huelsenbeck, J. P. and Ronquist, F. (2001). MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics, 17, 754–755.CrossRefGoogle ScholarPubMed
Huelsenbeck, J. P., Ronquist, F., Nielsen, R. and Bollback, J. P. (2001). Bayesian inference of phylogeny and its impact on evolutionary biology. Science, 294, 2310–2314.CrossRefGoogle ScholarPubMed
Hughes, B. D. (1995). Random Walks and Random Environments. Volume 1: Random Walks. Oxford: Clarendon Press.Google Scholar
Hur, K., Kim, J.-I., Choi, S.-I., et al. (2002). The pathogenic mechanisms of prion diseases. Mechanisms of Ageing and Development, 123, 1637–1647.CrossRefGoogle ScholarPubMed
Huse, G., Strand, E. and Giske, J. (1999). Implementing behaviour in individual-based models using neural networks and genetic algorithms. Evolutionary Ecology, 13, 469–483.CrossRefGoogle Scholar
Hutchings, J. A. (2000). Collapse and recovery of marine fishes. Nature, 406, 882–886.CrossRefGoogle ScholarPubMed
Hutchings, J. A. (2001). Influence of population decline, fishing, and spawner variability on the recovery of marine fishes. Journal of Fish Biology, 59 (Supplement A), 306–322.CrossRefGoogle Scholar
Hutchings, J. A. and Myers, R. A. (1994). What can be learned from the collapse of a renewable resource? Atlantic cod, Gadus morhua, of Newfoundland and Labrador. Canadian Journal of Fisheries and Aquatic Sciences, 51, 2126–2146.CrossRefGoogle Scholar
Hutchings, J. A., Walters, C. and Haedrich, R. L. (1997). Is scientific inquiry incompatible with government information control?Canadian Journal of Fisheries and Aquatic Sciences, 54, 1198.CrossRefGoogle Scholar
Iwasa, Y., Hakoyama, H., Nakamaru, M. and Nakanishi, J. (2000). Estimate of population extinction risk and its application to ecological risk management. Population Ecology, 42, 73–80.CrossRefGoogle Scholar
Jacobson, L. D. and MacCall, A. D. (1995). Stock–recruitment models for Pacific Sardine (Sardinops sagax). Canadian Journal of Fisheries and Aquatic Sciences, 52, 566–577.CrossRefGoogle Scholar
Jacobson, L. D., Oliveira, J. A. A., Barange, M., et al. (2001). Surplus production, variability, and climate change in the great sardine and anchovy fisheries. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1891–1903.CrossRefGoogle Scholar
Jacobson, L. D., Cadrin, S. X. and Weinberg, J. R. (2002). Tools for estimating surplus production and FMSY in any stock assessment model. North American Journal of Fisheries Management, 22, 326–338.2.0.CO;2>CrossRefGoogle Scholar
James, A., Pitchford, J. W. and Brindley, J. (2003). The relationship between plankton blooms, the hatching of fish larvae, and recruitment. Ecological Modelling, 160, 77–90.CrossRefGoogle Scholar
Janssen, A. (1989). Optimal host selection by Drosophila parasitoids in the field. Functional Ecology, 3, 469–479.CrossRefGoogle Scholar
Janssen, A., Driessen, G., Haan, M. and Roodbol, N. (1988). The impact of parasitoids on natural populations of temperature woodland Drosophila. Netherlands Journal of Zoology, 38, 61–73.CrossRefGoogle Scholar
Jaynes, E. T. (2003). Probability Theory. The Logic of Science. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Jenkins, R. (2001). Churchill. New York: Farrar, Straus and Grioux.Google Scholar
Jennings, S. (2000). Patterns and prediction of population recovery in marine reserves. Reviews in Fish Biology and Fisheries, 10, 209–231.CrossRefGoogle Scholar
Jennings, S. and Polunin, N. V. C. (1996). Fishing strategies, fishery development and socioeconomics in traditionally managed Fijian fishing grounds. Fisheries Management and Ecology, 3, 335–347.CrossRefGoogle Scholar
Jennings, S., Reynolds, J. D. and Mills, S. C. (1998). Life history correlates of responses to fisheries exploitation. Proceedings of the Royal Society of London, B265, 333–339.CrossRefGoogle Scholar
Jennings, S., Kaiser, M. and Reynolds, J. D. (2001). Marine Fisheries Ecology. Oxford: Blackwell Science.Google Scholar
Jervis, M. A. and Kidd, N. A. C. (1986). Host-feeding strategies in hymenopteran parasitoids. Biological Reviews, 61, 395–434.CrossRefGoogle Scholar
Jervis, M. A., Heimpel, G. E., Ferns, P. N., Harvey, J. A. and Kidd, N. A. C. (2001). Life-history strategies in parasitoid wasps: a comparative analysis of ‘ovigeny’. Journal of Animal Ecology, 70, 442–458.CrossRefGoogle Scholar
Jonzen, N., Cardinale, M., Gardmark, A., Arrhenius, F. and Lundberg, P. (2002). Risk of collapse in the eastern Baltic cod fishery. Marine Ecology – Progress Series, 240, 225–233.CrossRefGoogle Scholar
Kac, M. (1985). Enigmas of Chance. New York: Harper and Row.Google Scholar
Kailath, T. (1980). Linear Systems. Englewood Cliffs, NJ: Prentice Hall.Google Scholar
Kaplan, E. H. and Wein, L. M. (2003). Smallpox eradication in west and central Africa: surveillance-containment or herd immunity?Epidemiology, 14, 90–92.CrossRefGoogle ScholarPubMed
Kaplan, E. H., Craft, D. L. and Wein, L. M. (2002). Emergency response to a smallpox attack: the case for mass vaccination. Proceedings of the National Academy of Sciences, 99, 10 935–10 940.CrossRefGoogle ScholarPubMed
Karlin, S. and Taylor, H. M. (1981). A Second Course in Stochastic Processes. New York: Academic Press.Google Scholar
Keeling, M. J. and Grenfell, B. T. (1997). Disease extinction and community size: modeling the persistence of measles. Science, 275, 65–67.CrossRefGoogle ScholarPubMed
Keeling, M. J. and Grenfell, B. T. (2000). Individual-based perspectives on R0. Journal of Theoretical Biology, 203, 51–61.CrossRefGoogle Scholar
Kehler, D. G., Myers, R. A. and Field, C. A. (2002). Measurement error and bias in the maximum reproductive rate for the Ricker model. Canadian Journal of Fisheries and Aquatic Sciences, 59, 854–864.CrossRefGoogle Scholar
Keizer, J. E. (1987). Statistical Thermodynamics of Nonequilibrium Processes. New York: Springer Verlag.CrossRefGoogle Scholar
Keller, E. F. (2002). Making Sense of Life. Cambridge, MA: Harvard University Press.Google Scholar
Keller, J. B. (1974). Optimal velocity in a race. American Mathematical Monthly, 81, 474–480.CrossRefGoogle Scholar
Kendall, M. and Stuart, A. (1979). The Advanced Theory of Statistics. Volume 2. Inference and Relationship, 4th edn. London: Charles Griffin and Company Limited.Google Scholar
Kermack, W. O. and McKendrick, A. G. (1927). A contribution to the mathematical theory of epidemics. Proceedings of the Royal Society of London, A115, 700–721.CrossRefGoogle Scholar
Kermack, W. O. and McKendrick, A. G. (1932). Contributions to the mathematical theory of epidemics – II. The problem of endemicity. Proceedings of the Royal Society of London, A138, 55–83.CrossRefGoogle Scholar
Kermack, W. O. and McKendrick, A. G. (1933). Contributions to the mathematical study of epidemics. III. Further studies of the problem of endemicity. Proceedings of the Royal Society of London, A141, 94–122.CrossRefGoogle Scholar
Kimura, M. and Ohta, T. (1971). Theoretical Aspects of Population Genetics. Princeton, NJ: Princeton University Press.Google ScholarPubMed
Kinas, P. G. (1996). Bayesian fishery stock assessment and decision making using adaptive importance sampling. Canadian Journal of Fisheries and Aquatic Sciences, 53, 414–423.CrossRefGoogle Scholar
King, J. R. and McFarlane, G. A. (2003). Marine fish life history strategies: applications to fishery management. Fisheries Management and Ecology, 10, 249–264.CrossRefGoogle Scholar
King, M. (1995). Fisheries Biology, Assessment and Management. Ames, Iowa: Iowa State University Press.Google Scholar
King, S. (2000). On Writing. New York: Scribner.Google Scholar
Kingsland, S. E. (1985). Modeling Nature. Chicago: University of Chicago Press.Google Scholar
Kirchner, J. W. (2001). Fractal power spectra plotted upside-down. Comment on “Scaling of power spectrum of extinction events in the fossil record” by V. P. Dimri and M. R. Prakash. Earth and Planetary Science Letters, 192, 617–621.CrossRefGoogle Scholar
Kirchner, J. W. (2002). Evolutionary speed limits inferred from the fossil record. Nature, 415, 65–68.CrossRefGoogle ScholarPubMed
Kirchner, J. W. and Roy, B. A. (1999). The evolutionary advantages of dying young: epidemiological implications of longevity in metapopulations. American Naturalist, 154, 140–159.CrossRefGoogle ScholarPubMed
Kirchner, J. W. and Weil, A. (1998). No fractals in fossil extinction statistics. Nature, 395, 337–338.CrossRefGoogle Scholar
Kirchner, J. W. and Weil, A. (2000). Correlations in fossil extinction and origination rates through geological time. Proceedings of the Royal Society of London, B267, 1301–1309.CrossRefGoogle ScholarPubMed
Kirkwood, T. B. L. (1999). Evolution, molecular biology and mortality plateaus. In Bohr, V. A., Clark, B. F. C. and Stevensner, T. (editors) Molecular Biology of Aging, Alfred Benzon Symposium 44. Copenhagen: Munksgaard, pp. 383–390.Google Scholar
Klein, E. K., Lavigne, C., Foueillassar, X., Gouyon, P. and Larédo, C. (2003). Corn pollen dispersal: quasi-mechanistic models and field experiments. Ecological Monographs, 73, 131–150.CrossRefGoogle Scholar
Klein, G. (1952). Mean first-passage times of Brownian motion and related problems. Proceedings of the Royal Society of London, A211, 431–443.CrossRefGoogle Scholar
Klyashtorin, L. B. (1998). Long-term climate change and main commercial fish production in the Atlantic and Pacific. Fisheries Research, 37, 115–125.CrossRefGoogle Scholar
Knell, R. J., Begon, M. and Thompson, D. J. (1996). Transmission dynamics of Bacillus thuringiensis infecting Plodia interpunctella: a test of the mass action assumption with an insect pathogen. Proceedings of the Royal Society of London, B263, 75–81.CrossRefGoogle ScholarPubMed
Koella, J. C. and Boëte, C. (2003). A model for the coevolution of immunity and immune evasion in vector-borne diseases with implications for the epidemiology of malaria. The American Naturalist, 161, 698–707.CrossRefGoogle ScholarPubMed
Koella, J. C. and Restif, O. (2001). Coevolution of parasite virulence and host life history. Ecology Letters, 4, 207–214.CrossRefGoogle Scholar
Koeller, P. (2003). The lighter side of reference points. Fisheries Research, 62, 1–6.CrossRefGoogle Scholar
Kolata, G. B. (1977). Catastrophe theory: the emperor has no clothes. Science, 196, 287 + 350–351.CrossRefGoogle ScholarPubMed
Koopman, B. O. (1980). Search and Screening. General Principles with Historical Applications. Elmsford, NY: Pergamon Press.Google Scholar
Korman, J. and Higgins, P. S. (1997). Utility of escapement time series data for monitoring the response of salmon populations to habitat alteration. Canadian Journal of Fisheries and Aquatic Sciences, 54, 2058–2067.CrossRefGoogle Scholar
Kot, M. (2001). Elements of Mathematical Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Kot, M., Lewis, M. A. and Driessche, P. (1996). Dispersal data and the spread of invading organisms. Ecology, 77, 2027–2042.CrossRefGoogle Scholar
Kramers, H. A. (1940). Brownian motion in a field of force and the diffusion model of chemical reactions. Physica, 7, 284–312.CrossRefGoogle Scholar
Kubo, R., Matsuo, K. and Kitahara, K. (1973). Fluctuation and relaxation of macrovariables. Journal of Statistical Physics, 9, 51–96.CrossRefGoogle Scholar
Kuikka, S., Hilden, M., Gislason, H. F., et al. (1999). Modeling environmentally driven uncertainties in Baltic cod (Gadus morhua) management by Bayesian inference diagrams. Canadian Journal of Fisheries and Aquatic Sciences, 56, 629–641.CrossRefGoogle Scholar
Laakso, J., Kaitala, V. and Ranta, E. (2003). Non-linear biological responses to disturbance: consequences on population dynamics. Ecological Modelling, 162, 247–258.CrossRefGoogle Scholar
Lande, R. (1985). Expected time for random genetic drift of a population between stable phenotypic states. Proceedings of the National Academy of Sciences, 82, 7641–7645.CrossRefGoogle ScholarPubMed
Lande, R. (1987). Extinction thresholds in demographic models of territorial species. American Naturalist, 130, 624–635.CrossRefGoogle Scholar
Lande, R., Engen, S. and Saether, B.-E. (2003). Stochastic Population Dynamics in Ecology and Conservation. Oxford: Oxford University Press.CrossRefGoogle Scholar
Lander, A. D. (2004). A calculus of purpose. PLoS Biology, 2, 0712–0714.CrossRefGoogle ScholarPubMed
Laurenson, M. K., Norman, R. A., Gilbert, L., Reid, H. W. and Hudson, P. J. (2003). Identifying disease reservoirs in complex systems: mountain hares as reservoirs of ticks and louping-ill virus, pathogens of red grouse. Journal of Animal Ecology, 72, 177–185.CrossRefGoogle Scholar
Law, R. (2000). Fishing, selection, and phenotypic evolution. ICES Journal of Marine Science, 57, 659–668.CrossRefGoogle Scholar
Law, R. and Dicekmann, U. (1998). Symbiosis through exploitation and the merger of lineages in evolution. Proceedings of the Royal Society of London, B265, 1245–1253.CrossRefGoogle Scholar
Law, R., Murrell, D. J. and Dieckmann, U. (2003). Population growth in space and time: spatial logistic equations. Ecology, 84, 252–262.CrossRefGoogle Scholar
Leggett, W. C. and Deblois, E. (1994). Recruitment in marine fishes: is it regulated by starvation and predation in the egg and larval stages?Netherlands Journal of Sea Research, 32, 119–134.CrossRefGoogle Scholar
Lenski, R. E. and May, R. M. (1994). The evolution of virulence in parasites and pathogens: reconciliation between two competing hypotheses. Journal of Theoretical Biology, 169, 253–265.CrossRefGoogle ScholarPubMed
Leonard, T. and Hsu, J. S. J. (1999). Bayesian Methods. Cambridge: Cambridge University Press.Google Scholar
Levin, S. A. and Segel, L. A. (1985). Pattern generation in space and aspect. SIAM Review, 27, 45–67.CrossRefGoogle Scholar
Levins, R. (1966). The strategy of model building in population biology. American Scientist, 54, 421–431.Google Scholar
Levins, R. (1969). Some demographic and genetic consequences of environmental heterogeneity for biological control. Bulletin of the Entomological Society of America, 15, 237–240.CrossRefGoogle Scholar
Levins, R. (1970). Extinction. In Gerstenhaber, M. (editor) Some Mathematical Questions in Biology. Providence, RI: American Mathematical Society, pp. 75–107.Google Scholar
Lewin, R. (1986). Punctuated equilibrium is now old hat. Science, 231, 672–673.CrossRefGoogle ScholarPubMed
Lewis, W. J., and Takasu, K. (1990). Use of learned odours by a parasitic wasp in accordance with host and food needs. Nature, 348, 635–636.CrossRefGoogle Scholar
Lewontin, R. and Cohen, D. (1969). On population growth in a randomly varying environment. Proceedings of the National Academy of Sciences, 62, 1056–1060.CrossRefGoogle Scholar
Liermann, M. and Hilborn, R. (1997). Depensation in fish stocks: a hierarchic Bayesian meta-analysis. Canadian Journal of Fisheries and Aquatic Sciences, 54, 1976–1984.CrossRefGoogle Scholar
Lighthill, M. J. (1958). Introduction to Fourier Analysis and Generalised Functions. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lima, S. L. (2002). Putting predators back into behavioral predator-prey interactions. Trends in Ecology and Evolution, 17, 70–75.CrossRefGoogle Scholar
Lin, C. C. and Segel, L. A. (1988 (1974)). Mathematics Applied to Deterministic Problems in the Natural Sciences. Philadelphia: SIAM (Society for Industrial and Applied Mathematics).CrossRefGoogle Scholar
Lin, Z.-S. (2003). Simulating unintended effects restoration. Ecological Modelling, 164, 169–175.CrossRefGoogle Scholar
Lindholm, J. B., Auster, P. J., Ruth, M. and Kaufman, L. (2001). Modeling the effects of fishing and implications for the design of marine protected areas: juvenile fish responses to variations in seafloor habitat. Conservation Biology, 15, 424–437.CrossRefGoogle Scholar
Lindley, S. T. (2003). Estimation of population growth and extinction parameters from noisy data. Ecological Applications, 13, 806–813.CrossRefGoogle Scholar
Link, D. R., Natale, G., Shao, R., et al. (1997). Spontaneous formation of macroscopic chiral domains in a fluid smectic phase of achiral molecules. Science, 278, 1924–1927.CrossRefGoogle Scholar
Link, J. S., Brodziak, J. K. T., Edwards, S. F., et al. (2002). Marine ecosystem assessment in a fisheries management context. Canadian Journal of Fisheries and Aquatic Sciences, 59, 1429–1440.CrossRefGoogle Scholar
Lipp, E. K., Huq, A. and Colwell, R. K. (2002). Effects of global climate on infectious disease: the cholera model. Clinical Microbiology Reviews, 15, 757–770.CrossRefGoogle ScholarPubMed
Lochmiller, R. L. and Deerenberg, C. (2000). Trade-offs in evolutionary immunology: just what is the cost of immunity?Oikos, 88, 87–98.CrossRefGoogle Scholar
Lockwood, D. R., Hastings, A. M. and Botsford, L. W. (2002). The effects of dispersal patterns on marine reserves: does the tail wag the dog?Theoretical Population Biology, 61, 297–310.CrossRefGoogle ScholarPubMed
LoGiudice, K., Ostfeld, R. S., Schmidt, K. A. and Keesing, F. (2003). The ecology of infectious disease: effects of host diversity and community composition on Lyme disease risk. Proceedings of the National Academy of Science, 100, 567–571.CrossRefGoogle ScholarPubMed
Lorenzen, K. (2000). Population dynamics and management. In Beveridge, M. C. M. and McAndrew, B. J. (editors) Tilapias: Biology and Exploitation. Dordrecht: Kluwer Academic Publishers, pp. 163–226.CrossRefGoogle Scholar
Ludwig, D. (1975). Persistence of dynamical systems under random perturbations. SIAM Review, 17, 605–640.CrossRefGoogle Scholar
Ludwig, D. (1981). Escape from domains of attraction for systems perturbed by noise. In Enns, R. H., Jones, B. L., Miura, R. M. and Rangnekar, S. S. (editors) Nonlinear Phenomena in Physics and Biology. New York: Plenum Press.CrossRefGoogle Scholar
Ludwig, D. (1995). Uncertainty and fisheries management. Lecture Notes in Biomathematics, 100, 516–528.CrossRefGoogle Scholar
Ludwig, D. (1999). Is it meaningful to estimate a probability of extinction?Ecology, 80, 298–310.CrossRefGoogle Scholar
Ludwig, D., Mangel, M. and Haddad, B. (2001). Ecology, conservation and public policy. Annual Review of Ecology and Systematics, 32, 481–517.CrossRefGoogle Scholar
MacArthur, R. H. and Pianka, E. R. (1966). On the optimal use of a patchy environment. American Naturalist, 100, 603–609.CrossRefGoogle Scholar
MacArthur, R. H. and Wilson, E. O. (1967). The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
MacCall, A. D. (1990). Dynamic Geography of Marine Fish Populations. Seattle, WA: University of Washington Press.Google Scholar
MacCall, A. D. (1998). Use of decision tables to develop a precautionary approach to problems in behavior, life history and recruitment variability. In Restrepo, V. R. (editor) Proceedings of the Fifth NMFS Stock Assessment Workshop: Providing Scientific Advice to Implement the Precautionary Approach under the Magnuson–Stevens Fishery Conservation and Management Act. Key Largo, FL: US Department of Commerce.Google Scholar
MacCall, A. D. (2002). Use of known-biomass production models to determine productivity of West Coast groundfish stocks. North American Journal of Fisheries Management, 22, 272–279.2.0.CO;2>CrossRefGoogle Scholar
MacDonald, N. (1989). Biological Delay Systems: Linear Stability Theory. Cambridge: Cambridge University Press.Google Scholar
Mackauer, M. and Voelkl, W. (1993). Regulation of aphid populations by aphidiid wasps: does parasitoid foraging behaviour or hyperparasitism limit impact? Oecologia, 94, 339–350.CrossRefGoogle ScholarPubMed
Malloch, P. D. (1994). Life-History and Habits of the Salmon, Sea-Trout, Trout and Other Freshwater Fish. Derrydale Press.Google Scholar
Mangel, M. (1982). Applied mathematicians and naval operators. SIAM Review, 24, 289–300.CrossRefGoogle Scholar
Mangel, M. (1985). Decision and Control in Uncertain Resource Systems. New York: Academic Press.Google Scholar
Mangel, M. (1992). Descriptions of superparasitism by optimal foraging theory, evolutionary stable strategies, and quantitative genetics. Evolutionary Ecology, 6, 152–169.CrossRefGoogle Scholar
Mangel, M. (1998). No-take areas for sustainability of harvested species and a conservation invariant for marine reserves. Ecology Letters, 1, 87–90.CrossRefGoogle Scholar
Mangel, M. (2000a). Irreducible uncertainties, sustainable fisheries and marine reserves. Evolutionary Ecology Research, 2, 547–557.Google Scholar
Mangel, M. (2000b). On the fraction of habitat allocated to marine reserves. Ecology Letters, 3, 15–22.CrossRefGoogle Scholar
Mangel, M. (2000c). Trade-offs between fish habitat and fishing mortality and the role of reserves. Bulletin of Marine Science, 66, 663–674.Google Scholar
Mangel, M. (2001a). Complex adaptive systems, aging and longevity. Journal of Theoretical Biology, 213, 559–571.CrossRefGoogle Scholar
Mangel, M. (2001b). Required reading for (ecological) battles. (Review of N Eldredge. The Triumph of Evolution … and the Failure of Creationism). Trends in Ecology and Evolution, 16, 110.CrossRefGoogle Scholar
Mangel, M. and Beder, J. H. (1985). Search and stock depletion: theory and applications. Canadian Journal of Fisheries and Aquatic Sciences, 42, 150–163.CrossRefGoogle Scholar
Mangel, M. and Clark, C. W. (1988). Dynamic Modeling in Behavioral Ecology. Princeton, NJ: Princeton University Press.Google Scholar
Mangel, M. and Ludwig, D. (1977). Probability of extinction in a stochastic competition. SIAM Journal on Applied Mathematics, 33, 256–266.CrossRefGoogle Scholar
Mangel, M. and Roitberg, B. D. (1992). Behavioral stabilization of host–parasitoid population dynamics. Theoretical Population Biology, 42, 308–320.CrossRefGoogle Scholar
Mangel, M. and Tier, C. (1993). Dynamics of metapopulations with demographic stochasticity and environmental catastrophes. Theoretical Population Biology, 44, 1–31.CrossRefGoogle Scholar
Mangel, M. and Tier, C. (1994). Four facts every conservation biologist should know about persistence. Ecology, 75, 607–614.CrossRefGoogle Scholar
Mangel, M., Hofman, R. J., Norse, E. A. and Twiss, J. R. (1993). Sustainability and ecological research. Ecological Applications, 3, 573–575.Google ScholarPubMed
Mangel, M., Mullan, A., Mulch, A., Staub, S. and Yasukochi, E. (1998). A generally accessible derivation of the golden rule of bioeconomics. Bulletin of the Ecological Society of America, 79, 145–148.Google Scholar
Mangel, M., Fiksen, O. and Giske, J. (2001). Theoretical and statistical models in natural resource management and research. In Shenk, T. M. and Franklin, A. B. (editors) Modeling in Natural Resource Management. Development, Interpretation, and Application. Washington, DC: Island Press, pp. 57–72.Google Scholar
Mantua, N. J., Hare, S. R., Zhang, Y., Wallace, J. M. and Francis, R. C. (1997). A Pacific interdecadal climate oscillation with impacts on salmon production. Bulletin of the American Meteorological Society, 78, 1069–1079.2.0.CO;2>CrossRefGoogle Scholar
Margulis, L. and Sagan, D. (2002). Acquiring Genomes. A Theory of the Origins of Species. New York: Basic Books.Google Scholar
Marris, G., Hubbar, S. and Hughes, J. (1986). Use of patchy resources by Nemeritis canescens (Hymenoptera: Ichneumonidae). I. Optimal solutions. Journal of Animal Ecology, 55, 631–640.CrossRefGoogle Scholar
Marshall, C. T., Kjesbu, O. S., Yaragina, N. A., Solemdal, P. and Ulltang, O. (1998). Is spawner biomass a sensitive measure of the reproductive and recruitment potential of Northeast Arctic cod? Canadian Journal of Fisheries and Aquatic Sciences, 55, 1766–1783.CrossRefGoogle Scholar
Martz, H. F. and Waller, R. A. (1982). Bayesian Reliability Analysis. New York: John Wiley and Sons.Google Scholar
Masel, J. and Bergman, A. (2003). The evolution of the evolvability properties of the yeast prion ‘PSI + ’. Evolution, 57, 1498–1512.CrossRefGoogle Scholar
Matsuda, H. and Nishimori, K. (2003). A size-structured model for a stock-recovery program for an exploited endemic fisheries resource. Fisheries Research, 60, 223–236.CrossRefGoogle Scholar
Maunder, M. (2003). Is it time to discard the Schaefer model from the stock assessment scientist's toolbox? Fisheries Research, 61, 145–149.CrossRefGoogle Scholar
Maunder, M. N. (2002). The relationship between fishing methods, fisheries management and the estimation of maximum sustainable yield. Fish and Fisheries, 3, 251–260.CrossRefGoogle Scholar
May, R. M. (1974). Stability and Complexity in Model Ecosystems, 2nd edn. Princeton, NJ: Princeton University Press.Google Scholar
May, R. M. and Anderson, R. M. (1978). Regulation and stability of host-parasite population interactions. II. Destabilizing processes. Journal of Animal Ecology, 47, 249–267.CrossRefGoogle Scholar
May, R. M., Conway, G. R., Hassel, M. P. and Southwood, T. R. E. (1974). Time delays, density dependence and single species oscillations. Journal of Animal Ecology, 43, 747–770.CrossRefGoogle Scholar
Maynard Smith, J. (1968). Some Mathematical Ideas in Biology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Maynard Smith, J. (1982). Evolution and the Theory of Games. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
McAllister, M. K. (1996). Applications of Bayesian decision theory to fisheries policy formation: review. In Arnason, R. and Davidsson, T. B. (editors) Essays on Statistical Modelling Methodology for Fisheries Management. Rekjavek, Iceland: The Fisheries Research Institute University Press.Google Scholar
McAllister, M. K. and Ianelli, J. N. (1997). Bayesian stock assessment using catch-age data and the sampling-importance resampling algorithm. Canadian Journal of Fisheries and Aquatic Sciences, 54, 284–300.Google Scholar
McAllister, M. K. and Kirkwood, G. P. (1998a). Bayesian stock assessment: a review and example application using the logistic model. ICES Journal of Marine Science, 55, 1031–1060.CrossRefGoogle Scholar
McAllister, M. K. and Kirkwood, G. P. (1998b). Using Bayesian decision analysis to help achieve a precautionary approach for managing developing fisheries. Canadian Journal of Fisheries and Aquatic Sciences, 55, 2642–2661.CrossRefGoogle Scholar
McAllister, M. K. and Kirkwood, G. P. (1999). Applying multivariate conjugate priors in fishery-management system evaluation: how much quicker is it and does it bias the ranking of management options. ICES Journal of Marine Science, 56, 884.CrossRefGoogle Scholar
McAllister, M. K. and Peterman, R. M. (1992a). Decision analysis of a large-scale fishing experiment designed to test for a genetic effect of size-selective fishing on British Columbia pink salmon (Oncorhynchus gorbuscha). Canadian Journal of Fisheries and Aquatic Sciences, 49, 1305–1314.CrossRefGoogle Scholar
McAllister, M. K. and Peterman, R. M. (1992b). Experimental design in the management of fisheries: A review. North American Journal of Fisheries Management, 12, 1–18.2.3.CO;2>CrossRefGoogle Scholar
McAllister, M. K., Pikitch, E. K., Punt, A. E. and Hilborn, R. (1994). A Bayesian approach to stock assessment and harvest decision using the sampling/importance resampling algorithm. Canadian Journal of Fisheries and Aquatic Sciences, 51, 2673–2687.CrossRefGoogle Scholar
McAllister, M. K., Starr, P. J., Restrepo, V. R. and Kirkwood, G. P. (1999). Formulating quantitative methods to evaluate fishery-management systems: what fishery processes should be modelled and what trade-offs should be made? ICES Journal of Marine Science, 56, 900–916.CrossRefGoogle Scholar
McAllister, M. K., Pikitch, E. K. and Babcock, E. A. (2001). Using demographic methods to construct Bayesian priors for the intrinsic rate of increase in the Schaefer model and implications for stock rebuilding. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1871–1890.CrossRefGoogle Scholar
McCallum, H., Barlow, N. and Hone, J. (2001). How should pathogen transmission be modelled? Trends in Ecology and Evolution, 16, 295–300.CrossRefGoogle ScholarPubMed
McClanahan, T. R. and Kaunda-Arara, B. (1996). Fishery recovery in a coral-reef marine park and its effect on the adjacent fishery. Conservation Biology, 10, 1187–1199.CrossRefGoogle Scholar
McDonald, A. D., Sandal, L. K. and Steinshamn, S. I. (2002). Implications of a nested stochastic/deterministic bio-economic model for a pelagic fishery. Ecological Modelling, 149, 193–201.CrossRefGoogle Scholar
McDowall, R. M. (1988). Diadromy in Fishes. Migrations Between Freshwater and Marine Environments. Portland, OR: Timber Press.Google Scholar
McGarvey, R. (2003). Demand-side fishery management: integrating two forms of input control. Marine Policy, 27, 207–218.CrossRefGoogle Scholar
McGregor, R. (1997). Host-feeding and oviposition by parasitoids on hosts of different fitness value: influences of egg load and encounter rate. Journal of Insect Behavior, 10, 451–462.CrossRefGoogle Scholar
McGregor, R. R. and Roitberg, B. D. (2000). Size-selective oviposition by parasitoids and the evolution of life-history timing in hosts: fixed preference vs frequency-dependent host selection. Oikos, 89, 305–312.CrossRefGoogle Scholar
Mchich, R., Auger, P. M., Parra, R. B. and Raissi, N. (2002). Dynamics of a fishery on two fishing zones with fish stock dependent migrations: aggregation and control. Ecological Modelling, 158, 51–62.CrossRefGoogle Scholar
McLeod, P., Martin, A. P. and Richards, K. J. (2002). Minimum length scale for growth-limited oceanic plankton distributions. Ecological Modelling, 158, 111–120.CrossRefGoogle Scholar
McNamara, J. M., Houston, A. I. and Collins, E. J. (2001). Optimality models in behavioral biology. SIAM Review, 43, 413–466.CrossRefGoogle Scholar
McNeill, J. R. (2000). Something New Under the Sun: an Environmental History of the Twentieth-Century World. New York: Norton.Google Scholar
McPhee, J. (2002). The Founding Fish. New York: Farrar, Straus and Giroux.Google Scholar
McPhee, M. V. and Quinn, T. P. (1998). Factors affecting the duration of nest defense and reproductive lifespan of female sockeye salmon, Oncorhynchus nerka. Environmental Biology of Fishes, 51, 469–475.Google Scholar
Medvinsky, A. B., Petrovskii, S. V., Tikhonova, I. A., Malchow, H. and Li, B.-L. (2002). Spatiotemporal complexity of plankton and fish dynamics. SIAM Review, 44, 311–370.CrossRefGoogle Scholar
Meier, C., Senn, W., Hauser, R. and Zimmerman, M. (1994). Strange limits of stability in host-parasitoid systems. Journal of Mathematical Biology, 32, 563–572.CrossRefGoogle Scholar
Meltzer, M. (2003). Risks and benefits of preexposure and postexposure smallpox vaccination. Emerging Infectious Diseases, 9, 1363–1370.CrossRefGoogle ScholarPubMed
Meltzer, M. I., Damon, I., LeDuc, J. W. and Millar, D. J. (2001). Modeling potential responses to smallpox as a bioterrorist weapon. Emerging Infectious Diseases, 7, 959–968.CrossRefGoogle ScholarPubMed
Merton, R. C. (1971). Optimum consumption and portfolio rules in a continuous-time model. Journal of Economic Theory, 3, 373–413.CrossRefGoogle Scholar
Mesnil, B. (2003). The catch-survey analysis (CSA) method of fish stock assessment: an evaluation using simulated data. Fisheries Research, 63, 193–212.CrossRefGoogle Scholar
Meyer, R. and Millar, R. B. (1999a). Bayesian stock assessment using a state-space implementation of the delay difference model. Canadian Journal of Fisheries and Aquatic Sciences, 56, 37–52.Google Scholar
Meyer, R. and Millar, R. B. (1999b). BUGS in Bayesian stock assessments. Canadian Journal of Fisheries and Aquatic Sciences, 56, 1078–1086.CrossRefGoogle Scholar
Millar, R. B. (2002). Reference priors for Bayesian fisheries models. Canadian Journal of Fisheries and Aquatic Sciences, 59, 1492–1502.CrossRefGoogle Scholar
Millar, R. B. and Meyer, R. (2000). Bayesian state-space modeling of age-structured data: fitting a model is just the beginning. Canadian Journal of Fisheries and Aquatic Sciences, 57, 43–50.CrossRefGoogle Scholar
Mills, D. (1989). Ecology and Management of Atlantic Salmon. London: Chapman and Hall.Google Scholar
Mitchell, W. A. and Valone, T. J. (1990). The Optimization Research Program: studying adaptations by their function. The Quarterly Review of Biology, 65, 43–52.CrossRefGoogle Scholar
Moerland, T. S. (1995). Temperature: enzyme and organelle. In Hochackha, P. and Mommsen, T. P. (editors) Biochemistry and Molecular Biology of Fishes. Amsterdam: Elsevier Science B.V.Google Scholar
Moret, Y. and Schmid-Hempel, P. (2000). Survival for immunity: the price of immune system activation for bumblebee workers. Science, 290, 1166–1167.CrossRefGoogle Scholar
Morgan, E. R., Milner-Gulland, E. J., Torgerson, P. R. and Medley, G. F. (2004). Ruminating on complexity: macroparasites of wildlife and livestock. Trends in Ecology and Evolution, 19, 181–188.CrossRefGoogle ScholarPubMed
Morris, W. F., Mangel, M. and Adler, F. R. (1995). Mechanisms of pollen deposition by insect pollinators. Evolutionary Ecology, 9, 304–317.CrossRefGoogle Scholar
Morse, P. M. and Kimball, G. E. (1951). Methods of Operations Research. Technology Press of Massachusetts Institute of Technology; and Cambridge and New York: Wiley.Google Scholar
Mosmann, T. R. and Sad, S. (1996). The expanding universe of T-cell subsets: Th1, Th2 and more. Immunology Today, 17, 138–146.CrossRefGoogle ScholarPubMed
Mosquera, J. and Adler, F. R. (1998). Evolution of virulence: a unified framework for coinfection and superinfection. Journal of Theoretical Biology, 195, 293–313.CrossRefGoogle ScholarPubMed
Mosquera, J., Côté, M., Jennings, S. and Reynolds, J. D. (2000). Conservation benefits of marine reserves for fish populations. Animal Conservation, 4, 321–332.CrossRefGoogle Scholar
Mueller, L. D. and Rose, M. R. (1996). Evolutionary theory predicts late-life mortality plateaus. Proceedings of the National Academy of Sciences, 93, 15 249–15 253.CrossRefGoogle ScholarPubMed
Mullan, M. (1993). Webs and Scales. Physical and Ecological Processes in Marine Recruitment. Seattle, WA: University of Washington Press.Google Scholar
Murawski, S. A. (2000). Definitions of overfishing from an ecosystem perspective. ICES Journal of Marine Science, 57, 649–658.CrossRefGoogle Scholar
Murdoch, W. W. (1994). Population regulation in theory and practice. Ecology, 75, 271–287.CrossRefGoogle Scholar
Murdoch, W. W., Nisbet, R., Blythe, S. P., Gurney, W. S. C. and Reeve, J. D. (1987). An invulnerable age class and stability in delay-differential parasitoid–host models. American Naturalist, 129, 263–282.CrossRefGoogle Scholar
Murdoch, W. W., Briggs, C. J. and Nisbet, R. (2003). Consumer–Resource Dynamics. Princeton, NJ: Princeton University Press.Google Scholar
Murray, J. D. (1990). Turing's theory of morphogenesis – its influence on modelling biological pattern and form. Bulletin of Mathematical Biology, 52, 119–153.CrossRefGoogle Scholar
Murray, J. D. (2002). Mathematical Biology I: An Introduction. New York: Springer Verlag.Google Scholar
Murray, J. D. (2003). Mathematical Biology II: Spatial Models and Biomedical Applications. New York: Springer Verlag.Google Scholar
Myers, R. A., Mertz, G. and Fowlow, P. S. (1997a). Maximum population growth rates and recovery times for Atlantic cod, Gadus morhua. Fishery Bulletin, 95, 762–772.Google Scholar
Myers, R. A., Hutchings, J. A. and Barrowman, N. J. (1997b). Why do fish stocks collapse? The example of cod in Atlantic Canada. Ecological Applications, 7, 91–106.CrossRefGoogle Scholar
Myers, R. A., MacKenzie, B. R., Bowen, K. G. and Barrowman, N. J. (2001). What is the carrying capacity for fish in the ocean? A meta-analysis of population dynamics of North Atlantic cod. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1464–1476.CrossRefGoogle Scholar
Myers, R. A., Barrowman, N. J., Hilborn, R. and Kehler, D. G. (2002). Inferring Bayesian priors with limited direct data: Applications to risk analysis. North American Journal of Fisheries Management, 22, 351–364.2.0.CO;2>CrossRefGoogle Scholar
Nakken, O., Sandberg, P. and Steinshamn, S. I. (1996). Reference points for optimal fish stock management. Marine Policy, 20, 447–462.CrossRefGoogle Scholar
Narasimhan, T. N. (1999). Fourier's heat conduction equation: history, influence, and connections. Reviews of Geophysics, 37, 151–172.CrossRefGoogle Scholar
Nasell, I. (2002). Stochastic models of some endemic infections. Mathematical Biosciences, 179, 1–19.CrossRefGoogle ScholarPubMed
Needle, C. L. (2002). Recruitment models: diagnosis and prognosis. Reviews in Fish Biology and Fisheries, 11, 95–111.CrossRefGoogle Scholar
Nesse, R. M. and Williams, G. C. (1994). Why We Get Sick. New York: Vintage Books.Google Scholar
Nicholson, A. J. (1933). The balance of animal populations. Journal of Animal Ecology, 2, 131–178.CrossRefGoogle Scholar
Nicholson, A. J. (1954). An outline of the dynamics of animal populations. Australian Journal of Zoology, 2, 9–65.CrossRefGoogle Scholar
Nicholson, A. J. and Bailey, V. A. (1935). The balance of animal populations. Proceedings of the Zoological Society of London, 3, 551–598.CrossRefGoogle Scholar
Nisbet, R. and Gurney, W. S. C. (1982). Modelling Fluctuating Populations. New York: John Wiley and Sons.Google Scholar
Nisbet, R. M. and Gurney, W. S. C. (1983). The systematic formulation of population models for insects with dynamically varying instar duration. Theoretical Population Biology, 23, 114–135.CrossRefGoogle Scholar
Nisbet, R. M., Blythe, S. P., Gurney, W. S. C. and Metz, J. A. J. (1985). Stage-structure models of populations with distinct growth and development processes. IMA Journal of Mathematics Applied in Medicine and Biology, 2, 57–68.CrossRefGoogle ScholarPubMed
Norse, E. A. (editor) (1993). Global Marine Biological Diversity. Washington, DC: Island Press.Google Scholar
Nowak, M. A. and May, R. M. (2000). Virus Dynamics. Mathematical Principles of Immunology and Virology. Oxford: Oxford University Press.Google Scholar
Nuland, S. B. (1993). How We Die. New York: Vintage Books.Google Scholar
O'Neill, P. D. (2002). A tutorial introduction to Bayesian inference for stochastic epidemic models using Markov chain Monte Carlo methods. Mathematical Biosciences, 180, 103–114.CrossRefGoogle ScholarPubMed
Oaks, S. C. Jr., Mitchell, V. S., Pearson, G. W. and Carpenter, C. C. J. (editors) (1991). Malaria. Obstacles and Opportunities. Washington, DC: National Academy Press.Google Scholar
Okey, T. A. (2003). Membership of the eight Regional Fishery Management Councils in the United States: are special interests over-represented? Marine Policy, 27, 193–206.CrossRefGoogle Scholar
Olson, D. M. and Andow, D. A. (1997). Primary sex allocation in Trichogramma (Hymenoptera: Trichogrammatidae) and the effects of sperm on oviposition behavior. Annals of the Entomological Society of America, 90, 689–692.CrossRefGoogle Scholar
Olson, D. M., Fadamiro, H., Lundgren, J. G. and Heimpel, G. E. (2000). Effects of sugar feeding on carbohydrate and lipid metabolism in a parasitoid wasp. Physiological Entomology, 25, 17–26.CrossRefGoogle Scholar
Olver, C., Shuter, B. J. and Minns, C. K. (1995). Toward a definition of conservation principles for fisheries management. Canadian Journal of Fisheries and Aquatic Sciences, 52, 1584–1594.CrossRefGoogle Scholar
Overholtz, W. J. (1999). Precision and uses of biological reference points calculated from stock recruitment data. North American Journal of Fisheries Management, 19, 643–657.2.0.CO;2>CrossRefGoogle Scholar
Overholtz, W. J., Edwards, S. F. and Brodziak, J. K. T. (1995). Effort control in the New England groundfish fishery: a bioeconomic perspective. Canadian Journal of Fisheries and Aquatic Sciences, 52, 1944–1957.CrossRefGoogle Scholar
Owen-Smith, N. (2002). Adaptive Herbivore Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Pacala, S. W., Hassell, M. P. and May, R. M. (1990). Host–parasitoid associations in patchy environments. Nature, 344, 150–153.CrossRefGoogle ScholarPubMed
Packer, C., Holt, R., Hudson, P., Lafferty, K. and Dobson, A. (2003). Keeping the herds healthy and alert: implications of predator control for infectious disease. Ecology Letters, 6, 797–802.CrossRefGoogle Scholar
Paddack, M. J. and Estes, J. A. (2000). Kelp forest fish populations in marine reserves and adjacent exploited areas of Central California. Ecological Applications, 10, 855–870.CrossRefGoogle Scholar
Pandolfini, B. (1989). Chess Openings: Traps and Zaps. New York: Simon and Schuster.Google Scholar
Patterson, K., Cook, R., Darby, C., et al. (2001). Estimating uncertainty in fish stock assessment and forecasting. Fish and Fisheries, 2, 125–157.CrossRefGoogle Scholar
Patterson, K. R. (1998). Assessing fish stocks when catches are misreported: model, simulation tests, and application to cod, haddock and whiting in the ICES area. ICES Journal of Marine Science, 55, 878–891.CrossRefGoogle Scholar
Patterson, K. R. (1999). Evaluating uncertainty in harvest control law catches using Bayesian Markov chain Monte Carlo virtual population analysis with adaptive rejection sampling and including structural uncertainty. Canadian Journal of Fisheries and Aquatic Sciences, 56, 208–221.CrossRefGoogle Scholar
Pearcy, W. G. (1992). Ocean Ecology of Pacific Salmonids. Seattle, WA: University of Washington Press.Google Scholar
Pearl, R. (1928). The Rate of Living. New York: Alfred Knopf.Google Scholar
Pearl, R. and Miner, J. R. (1935). Experimental studies on the duration of life. XIV. The comparative mortality of certain lower organisms. Quarterly Review of Biology, 10, 60–79.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1921). Experimental studies on the duration of life. I. Introductory discussion of the duration of life in drosophila. American Naturalist, 55, 481–509.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1922a). Experimental studies on the duration of life. II. Hereditary differences in duration of life in the line-bred strains of drosophila. American Naturalist, 56, 174–187.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1922b). Experimental studies on the duration of life. III. The effect of successive etherizations on the duration of life in drosophila. American Naturalist, 56, 273–280.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1922c). Experimental studies on the duration of life. IV. Data on the influence of density of population on duration of life in drosophila. American Naturalist, 56, 312–321.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1922d). Experimental studies on the duration of life. V. On the influence of certain environmental factors on the duration of life in drosophila. American Naturalist, 56, 385–405.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1924a). Experimental studies on the duration of life. IX. New life tables for drosophila. American Naturalist, 58, 71–82.CrossRefGoogle Scholar
Pearl, R. and Parker, S. A. (1924b). Experimental studies on the duration of life. X. The duration of life of drosophila melanogaster in the complete absence of food. American Naturalist, 58, 193–218.CrossRefGoogle Scholar
Pearl, R., Miner, J. R. and Parker, S. A. (1927). Experimental studies on the duration of life. XI. Density of population and life duration in drosophila. American Naturalist, 56, 289–318.CrossRefGoogle Scholar
Pearl, R., Parker, S. A. and Gonzalez, B. M. (1923). Experimental studies on the duration of life. VII. The Mendelian inheritance of duration of life in crosses of wild type and quintuple stocks of drosophila melanogaster. American Naturalist, 57, 153–192.CrossRefGoogle Scholar
Pearl, R., Parker, T. and Miner, J. R. (1941). Experimental studies on the duration of life. XVI. Life tables for the flour beetle Tribolium confusum duval. American Naturalist, 75, 5–19.CrossRefGoogle Scholar
Pellmyr, O. (2003). Yuccas, yucca moths, and coevolution: a review. Annals of the Missouri Botanical Garden, 90, 35–55.CrossRefGoogle Scholar
Perelson, A. S. and Nelson, P. W. (1999). Mathematical analysis of HIV-1 dynamics in vivo. SIAM Review, 41, 3–44.CrossRefGoogle Scholar
Perkins, P. C. and Edwards, E. F. (1995). A mixture model for estimating discarded bycatch from data with many zero observations: tuna discards in the eastern tropical Pacific Ocean. Fishery Bulletin, 94, 330–340.Google Scholar
Petchey, O. L. (2000). Environmental colour affects aspects of single-species population dynamics. Proceedings of the Royal Society of London, B267, 747–754.CrossRefGoogle ScholarPubMed
Petchey, O. L., Gonzalez, A. and Wilson, H. B. (1997). Effects on population persistence: the interaction between environmental noise colour, intraspecific competition and space. Proceedings of the Royal Society of London, B264, 1841–1847.CrossRefGoogle Scholar
Peterman, R. M. (1989). Application of statistical power analysis to the Oregon coho salmon (Oncorhynchus kisutch) problem. Canadian Journal of Fisheries and Aquatic Sciences, 46, 1183–1187.CrossRefGoogle Scholar
Peterman, R. M. (1990a). The importance of reporting statistical power: the forest decline and acid deposition example. Ecology, 71, 2024–2027.CrossRefGoogle Scholar
Peterman, R. M. (1990b). Statistical power analysis can improve fisheries research and management. Canadian Journal of Fisheries and Aquatic Sciences, 47, 2–15.CrossRefGoogle Scholar
Peters, C. S., Mangel, M. and Costantino, R. F. (1989). Stationary distribution of population size in Tribolium. Bulletin of Mathematical Biology, 51, 625–638.Google Scholar
Peters, R. H. (1983). The Ecological Implications of Body Size. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Peters, R. H. (1991). A Critique for Ecology. Cambridge: Cambridge University Press.Google Scholar
Pezzey, J. C. V., Roberts, C. M. and Urdal, B. T. (2000). A simple bioeconomic model of a marine reserve. Ecological Economics, 33, 77–91.CrossRefGoogle Scholar
Pielke, R. A. Jr. and Conant, R. T. (2003). Best practices in prediction for decision-making: lessons from the atmospheric and earth sciences. Ecology, 84, 1351–1358.CrossRefGoogle Scholar
Pierce, N. E. and Nash, D. R. (1999). The Imperial Blue: Jalmenus evagoras (Lycaenidae). In Kitching, R. L., Jones, R. E., Pierce, N. E., and Scheermeyer, E., eds., Biology of Australian Butterflies. Victoria: CSIRO Publishing, pp. 277–313.Google Scholar
Pierce, N. E., Braby, M. F., Heath, A., et al. (2002). The ecology and evolution of ant association in the Lycaenidae (Lepidoptera). Annual Review of Entomology, 47, 733–771.CrossRefGoogle Scholar
Pigliucci, M. and Murren, C. J. (2003). Genetical assimilation and a possible evolutionary paradox: can macroevolution sometimes be so fast as to pass us by? Evolution, 57, 1455–1464.CrossRefGoogle Scholar
Pikitch, E. K., Huppert, D. D. and Sissenwine, M. P. (editors) (1997). Global Trends: Fishery Management. Bethesda, MD: American Fisheries Society.Google Scholar
Pikitch, E. K., Santora, C., Babcock, E. A., et al. (2004). Ecosystem-based fishery management. Science, 305, 346–347.CrossRefGoogle ScholarPubMed
Pincock, R. E. and Wilson, K. R. (1973). Spontaneous generation of optical activity. Journal of Chemical Education, 50, 455–457.CrossRefGoogle Scholar
Pincock, R. E., Perkins, R. R., Ma, A. S. and Wilson, K. R. (1971). Probability distribution of enantiomorphous forms in spontaneous generation of optically active substances. Science, 174, 1018–1020.CrossRefGoogle ScholarPubMed
Pirsig, R. M. (1974). Zen and the Art of Motorcycle Maintenance. New York: William Morrow.Google Scholar
Pitcher, T. J. (2000). Ecosystem goals can reinvigorate fisheries management, help dispute resolution and encourage public support. Fish and Fisheries, 1, 99–103.CrossRefGoogle Scholar
Pitcher, T. J., Watson, R., Forrest, R., Valtysson, H. P. and Guénette, S. (2002). Estimating illegal and unreported catches from marine ecosystems: a basis for change. Fish and Fisheries, 3, 317–339.CrossRefGoogle Scholar
Pletcher, S. D. and Curtsinger, J. W. (1998). Mortality plateaus and the evolution of senescence: why are old-age mortality rates so low? Evolution, 52, 454–464.CrossRefGoogle ScholarPubMed
Prager, M. H. (2002). Comparison of logistic and generalized surplus-production models applied to swordfish, Xiphias gladius, in the north Atlantic Ocean. Fisheries Research, 58, 41–57.CrossRefGoogle Scholar
Prager, M. H., Porch, C. E., Shertzer, K. W. and Caddy, J. F. (2003). Targets and limits for management of fisheries: a simple probability-based approach. North American Journal of Fisheries Management, 23, 349–361.2.0.CO;2>CrossRefGoogle Scholar
Preston, S. H., Heuveline, P. and Guillot, M. (2001). Demography. Measuring and Modeling Population Processes. Oxford: Blackwell.Google Scholar
Pritchard, G. (1969). The ecology of a natural population of Queensland fruit fly, Dacus tryoni. Australian Journal of Zoology, 17, 293–311.CrossRefGoogle Scholar
Pritchett, V. S. (1990a). The Complete Collected Stories. New York: Random House.Google Scholar
Pritchett, V. S. (1990b). The Complete Collected Essays. New York: Random House.Google Scholar
Provine, W. B. (1986). Sewall Wright and Evolutionary Biology. Chicago, IL: University of Chicago Press.Google Scholar
Punt, A. E. and Hilborn, R. (1997). Fisheries stock assessment and decision analysis: the Bayesian approach. Reviews in Fish Biology and Fisheries, 7, 35–63.CrossRefGoogle Scholar
Puterman, M. L. (1994). Markov Decision Processes. Discrete Stochastic Dynamic Programming. New York: John Wiley and Sons.Google Scholar
Pybus, O., Charleston, M. A., Gupta, S., et al. (2001). The epidemic behavior of the hepatitis C virus. Science, 292, 2323–2325.CrossRefGoogle ScholarPubMed
Quinn, T. J. I. and Deriso, R. B. (1999). Quantitative Fish Dynamics. New York: Oxford University Press.Google Scholar
Raftery, A. E. (1988). Inference for the binomial N-parameter: A hierarchical Bayes approach. Biometrika, 75, 223–228.CrossRefGoogle Scholar
Railsback, S. F. (2001). Concepts from complex adaptive systems as a framework for individual-based modelling. Ecological Modelling, 139, 47–62.CrossRefGoogle Scholar
Read, A. F. (2003). Simplicity and serenity in advanced statistics. Trends in Ecology and Evolution, 18, 11–12.CrossRefGoogle Scholar
Read, A. F., Anwar, M., Shutler, D. and Nee, S. (1995). Sex allocation and population structure in malaria and related parasitic protozoa. Proceedings of the Royal Society of London, B260, 359–363.CrossRefGoogle ScholarPubMed
Read, A. F., Narara, A., Nee, S., Keymer, A. E. and Day, K. P. (1992). Gametocyte sex ratios as indirect measures of outcrossing rates in malaria. Parasitology, 104, 387–395.CrossRefGoogle ScholarPubMed
Reed, W. J. and Simons, C. M. (1996). Analyzing catch-effort data by means of the Kalman filter. Canadian Journal of Fisheries and Aquatic Sciences, 53, 2157–2166.Google Scholar
Rees, M., Mangel, M., Turnbull, L., Sheppard, A. and Briese, D. (2000). The effects of heterogeneity on dispersal and colonization in plants. In Hutchings, M. J., John, E. A., and Stewart, A. J. A., eds., The Ecological Consequences of Environmental Heterogeneity. Oxford: Blackwell Science, pp. 237–265.Google Scholar
Reid, C. (1976). Courant in Göttingen and New York. The Story of an Improbable Mathematician. New York: Springer Verlag.Google Scholar
Reidl, J. and Klose, K. E. (2002). Vibrio cholerae and cholera: out of the water and into the host. FEMS Microbiology Reviews, 26, 125–139.CrossRefGoogle Scholar
Ricciardi, L. M. and Sato, S. (1990). Diffusion processes and first-passage-time problems. In Ricciardi, L. M., ed., Lectures in Applied Mathematics and Informatics. New York: St. Martin's Press, pp. 206–284.Google Scholar
Richards, L. J. and Maguire, J.-J. (1998). Recent international agreements and the precautionary approach: new directions for fisheries management science. Canadian Journal of Fisheries and Aquatic Sciences, 55, 1545–1552.CrossRefGoogle Scholar
Rickman, S. J., Dulvy, N. K., Jennings, S. and Reynolds, J. D. (2000). Recruitment variation related to fecundity in marine fishes. Canadian Journal of Fisheries and Aquatic Sciences, 57, 116–124.CrossRefGoogle Scholar
Rivot, E., Prevost, E. and Parent, E. (2001). How robust are Bayesian posterior inferences based on a Ricker model with regards to measurement errors and prior assumptions about parameters? Canadian Journal of Fisheries and Aquatic Sciences, 58, 2284–2297.CrossRefGoogle Scholar
Robb, C. A. and Peterman, R. M. (1998). Application of Bayesian decision analysis to management of a sockeye salmon (Oncorhynchus nerka) fishery. Canadian Journal of Fisheries and Aquatic Sciences, 55, 86–98.CrossRefGoogle Scholar
Robertson, I. C., Robertson, W. G. and Roitberg, B. D. (1998). A model of mutual tolerance and the origin of communal associations between unrelated females. Journal of Insect Behavior, 11, 265–286.CrossRefGoogle Scholar
Roff, D. A. (1984). The evolution of life history parameters in teleosts. Canadian Journal of Fisheries and Aquatic Sciences, 41, 989–999.CrossRefGoogle Scholar
Roff, D. A. (1991). The evolution of life-history variation in fishes, with particular reference to flatfishes. Netherlands Journal of Sea Research, 27, 197–207.CrossRefGoogle Scholar
Rohani, P., Keeling, M. J. and Grenfell, B. T. (2002). The interplay between determinism and stochasticity in childhood diseases. American Naturalist, 159, 469–481.CrossRefGoogle ScholarPubMed
Roitberg, B. D. (1990). Optimistic and pessimistic fruit flies: evaluating fitness consequences of estimation errors. Behaviour, 114, 65–82.CrossRefGoogle Scholar
Roitberg, B. D. and Friend, W. G. (1992). A general theory for host seeking decisions in mosquitoes. Bulletin of Mathematical Biology, 54, 401–412.CrossRefGoogle ScholarPubMed
Roitberg, B. D. and Lalonde, R. G. (1991). Host marking enhance parasitism risk for a fruit-infesting fly Rhagoletis basiola. Oikos, 61, 389–393.CrossRefGoogle Scholar
Roitberg, B. D. and Prokopy, R. J. (1987). Insects that mark hosts. BioScience, 37, 400–406.CrossRefGoogle Scholar
Roitberg, B. D., Cailrl, R. S. and Prokopy, R. J. (1984). Oviposition deterring pheromone influences dispersal distance in tephritid fruit flies. Entomologia Experimentalis et Applicata, 35, 217–220.CrossRefGoogle Scholar
Roitberg, B. D., Mangel, M., Lalonde, R., et al. (1992). Seasonal dynamic shifts in patch exploitation by parasitic wasps. Behavioral Ecology, 3, 156–165.CrossRefGoogle Scholar
Roitberg, B. D., Sircom, J., Roitberg, C. A., Alphen, J. J. M. and Mangel, M. (1993). Life expectancy and reproduction. Nature, 364, 351.CrossRefGoogle ScholarPubMed
Rolff, J. and Siva-Jothy, M. T. (2002). Copulation corrupts immunity: a mechanism for a cost of mating in insects. PNAS, 99, 9916–9918.CrossRefGoogle ScholarPubMed
Rolff, J. and Siva-Jothy, M. T. (2003). Invertebrate ecological immunology. Science, 301, 472–475.CrossRefGoogle ScholarPubMed
Romagnani, S. (1996). TH1 and TH2 in human diseases. Clinical Immunology and Immunopathology, 80, 225–235.CrossRefGoogle ScholarPubMed
Root, K. V. (1998). Evaluating the effects of habitat quality, connectivity, and catastrophes on threatened species. Ecological Applications, 8, 854–865.CrossRefGoogle Scholar
Rosà, R., Pugliese, A., Villani, A. and Rizzoli, A. (2003). Individual-based vs. deterministic models for macroparasites: host cycles and extinction. Theoretical Population Biology, 63, 295–307.CrossRefGoogle ScholarPubMed
Rosenberg, A. A. (2003). Managing to the margins: the overexploitation of fisheries. Frontiers in Ecology and the Environment, 1, 102–106.CrossRefGoogle Scholar
Rosenberg, A. A. and Restrepo, V. R. (1994). Uncertainty and risk evaluation in stock assessment advice for U.S. marine fisheries. Canadian Journal of Fisheries and Aquatic Sciences, 51, 2715–2720.CrossRefGoogle Scholar
Rosenheim, J. A. (1996). An evolutionary argument for egg limitation. Evolution, 50, 2089–2094.CrossRefGoogle ScholarPubMed
Rosenheim, J. A. (1999). Characterizing the cost of oviposition in insects: a dynamic model. Evolutionary Ecology, 13, 141–165.CrossRefGoogle Scholar
Rosenheim, J. A. and Heimpel, G. E. (1998). Egg limitation in parasitoids: A review of the evidence and a case study. Biological Control, 11, 160–168.Google Scholar
Rosenheim, J. A. and Heimpel, G. E. (2000). Egg maturation, egg resorption, and the costliness of transient egg limitation in insects. Proceedings of the Royal Society of London, B267, 1565–1573.CrossRefGoogle ScholarPubMed
Rosenheim, J. A. and Mangel, M. (1994). Patch-leaving rules for parasitoids with imperfect host discrimination. Ecological Entomology, 19, 374–380.CrossRefGoogle Scholar
Rosenheim, J. A. and Rosen, D. (1992). Influence of egg load and host size on host-feeding behaviour of the parasitoid Aphytis lingnaensis. Ecological Entomology, 17, 263–272.CrossRefGoogle Scholar
Rosenzweig, M. L. (1995). Species Diversity in Space and Time. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Ross, R. (1916). An application of the theory of probabilities to the study of a priori pathometry. Part I. Proceedings of the Royal Society of London, A92, 204–230.CrossRefGoogle Scholar
Ross, R. and Hudson, H. P. (1917a). An application of the theory of probabilities to the study of a priori pathometry. Part II. Proceedings of the Royal Society of London, A93, 212–225.CrossRefGoogle Scholar
Ross, R. and Hudson, H. P. (1917b). An application of the theory of probabilities to the study of a priori pathometry. Part III. Proceedings of the Royal Society of London, A93, 225–240.CrossRefGoogle Scholar
Royall, R. (1997). Statistical Evidence. A Likelihood Paradigm. London: Chapman and Hall.Google Scholar
Rudd, M. A., Tupper, M. H., Folmer, H. and Kooten, G. C. (2003). Policy analysis for tropical marine reserves: challenges and directions. Fish and Fisheries, 4, 65–85.CrossRefGoogle Scholar
Ruxton, G. D. and Rohani, P. (1996). The consequences of stochasticity for self-organized spatial dynamics, persistence and coexistence in spatially extended host-parasitoid communities. Proceedings of the Royal Society of London, B263, 625–631.CrossRefGoogle Scholar
Sadovy, Y. (2001). The threat of fishing to highly fecund fishes. Journal of Fish Biology, 59 (Supplement A), 90–108.CrossRefGoogle Scholar
Saether, B.-E., Engen, S., Lande, R., Both, C. and Visser, M. E. (2002). Density dependence and stochastic variation in a newly established population of a small songbird. Oikos, 99, 331–337.CrossRefGoogle Scholar
Saether, B. E. and Engen, S. (2004). Stochastic population theory faces reality in the laboratory. Trends in Ecology and Evolution, 19, 351–353.CrossRefGoogle ScholarPubMed
Sanchez Lizaso, J. L., Goni, R., Renones, O., et al. (2000). Density dependence in marine protected populations: a review. Environmental Conservation, 27, 144–158.CrossRefGoogle Scholar
Sanchirico, J. N., Stoffle, R., Broad, K. and Talaue-McManus, L. (2003). Modeling marine protected areas. Science, 301, 47–48.CrossRefGoogle ScholarPubMed
Sauvage, F., Langlais, M., Yoccoz, N. G. and Pontier, D. (2003). Modelling hantavirus in fluctuating populations of bank voles: the role of indirect transmission on virus persistence. Journal of Animal Ecology, 72, 1–13.CrossRefGoogle Scholar
Scheffer, M. and Carpenter, S. (2003). Catastrophic regime shifts in ecosystems: linking theory to observation. Trends in Ecology and Evolution, 18, 648–656.CrossRefGoogle Scholar
Scheffer, M., Carpenter, S., Foley, J. A., Folke, C. and Walker, B. (2001). Catastrophic shifts in ecosystems. Nature, 413, 591–596.CrossRefGoogle ScholarPubMed
Schemske, D. W. and Bierzychudek, P. (2001). Perspective: evolution of flower color in the desert annual Linanthus parryae: Wright revisited. Evolution, 55, 1269–1282.CrossRefGoogle ScholarPubMed
Schluter, D. (2000). The Ecology of Adaptive Radiation. New York: Oxford University Press.Google Scholar
Schmid-Hempel, P. and Ebert, D. (2003). On the evolutionary ecology of specific immune defense. Trends in Ecology and Evolution, 18, 27–32.CrossRefGoogle Scholar
Schnute, J., Cass, A. and Richards, L. J. (2000). A Bayesian decision analysis to set escapement goals for Fraser River sockeye salmon (Oncorhynchus nerka). Canadian Journal of Fisheries and Aquatic Sciences, 57, 962–979.CrossRefGoogle Scholar
Schnute, J. T. (1993). Ambiguous inferences from fisheries data. In Barnett, V. and Turkman, K. F. (editors) Statistics for the Environment. London: John Wiley and Sons, pp. 293–309.Google Scholar
Schnute, J. T. and Kronlund, A. R. (1996). A management oriented approach to stock recruitment analysis. Canadian Journal of Fisheries and Aquatic Sciences, 53, 1281–1293.CrossRefGoogle Scholar
Schnute, J. T. and Kronlund, A. R. (2002). Estimating salmon stock-recruitment relationships from catch and escapement data. Canadian Journal of Fisheries and Aquatic Sciences, 59, 433–449.CrossRefGoogle Scholar
Schnute, J. and Richards, L. J. (1998). Analytical models for fishery reference points. Canadian Journal of Fisheries and Aquatic Sciences, 55, 515–528.CrossRefGoogle Scholar
Schnute, J. T. and Richards, L. J. (2001). Use and abuse of fishery models. Canadian Journal of Fisheries and Aquatic Sciences, 58, 10–17.CrossRefGoogle Scholar
Schrag, S. J. and Wiener, A. P. (1995). Emerging infectious disease: what are the relative roles of ecology and evolution? Trends in Ecology and Evolution, 10, 319–324.CrossRefGoogle ScholarPubMed
Schulman, L. S. (1981). Techniques and Applications of Path Integration. New York: Wiley Interscience.Google Scholar
Schuss, Z. (1980). Theory and Applications of Stochastic Differential Equations. New York: Wiley.Google Scholar
Schwartz, M. (2003). How the Cows Turned Mad. Berkeley, CA: University of California Press.Google Scholar
Schwarz, G. (1978). Estimating the dimension of a model. Annals of Statistics, 6, 461–464.CrossRefGoogle Scholar
Seber, G. A. F. (1982). The Estimation of Animal Abundance, 2nd edn. New York: MacMillan.Google Scholar
Settle, W. H. and Wilson, L. T. (1990). Invasion by the variegated leafhopper and biotic interactions: parasitism, competition, and apparent competition. Ecology, 71, 1461–1470.CrossRefGoogle Scholar
Sevenster, J. A., Ellers, J. and Driessen, G. (1998). An evolutionary argument for time limitation. Evolution, 52, 1241–1244.CrossRefGoogle ScholarPubMed
Shaw, D. J. and Dobson, A. P. (1995). Patterns of macroparasite abundance and aggregation in wildlife populations: a quantitative review. Parasitology, 111, S111–S113.CrossRefGoogle ScholarPubMed
Shears, N. T. and Babcock, R. C. (2003). Continuing trophic cascade effects after 25 years of no-take marine reserve protection. Marine Ecology – Progress Series, 246, 1–16.CrossRefGoogle Scholar
Sheldon, B. C. and Verhulst, S. (1996). Ecological immunology: costly parasite defences and trade-offs in evolutionary ecology. TREE, 11, 317–321.Google ScholarPubMed
Shelton, P. A. and Healey, B. P. (1999). Should depensation be dismissed as a possible explanation for the lack of recovery of the northern cod (Gadus morhua) stock? Canadian Journal of Fisheries and Aquatic Sciences, 56, 1521–1524.CrossRefGoogle Scholar
Shepherd, J. G. (1982). A versatile new stock recruitment relationship for fisheries, and the construction of sustainable yield curves. J. Conseil International Exploration de la Mer, 40, 67–75.CrossRefGoogle Scholar
Shepherd, J. G. (1999). Extended survivors analysis: An improved method for the analysis of catch-at-age data and abundance indices. ICES Journal of Marine Science, 56, 584–591.CrossRefGoogle Scholar
Sherman, K. and Alexander, L. M. (editors) (1989). Biomass Yields and Geography of Large Marine Ecosystems. Boulder, CO: Westview Press.Google Scholar
Sherman, K., Alexander, L. M. and Gold, B. D. (editors) (1990). Large Marine Ecosystems. Patterns, Process and Yields. Washington, DC: American Association for the Advancement of Science.Google Scholar
Sherman, K., Alexander, L. M. and Gold, B. D. (editors) (1991). Food Chains, Yields, Models and Management of Large Marine Ecosystems. Boulder, CO: Westview Press.Google Scholar
Sherman, K., Alexander, L. M. and Gold, B. D. (editors) (1993). Large Marine Ecosystems. Stress, Mitigation and Sustainability. Washington, DC: American Association for the Advancement of Science.Google Scholar
Shudo, E. and Iwasa, Y. (2001). Inducible defense against pathogens and parasites: optimal choice among multiple options. Journal of Theoretical Biology, 209, 233–247.CrossRefGoogle ScholarPubMed
Siegel, J. S. (2002). Shattered mirrors. Nature, 419, 346–347.CrossRefGoogle ScholarPubMed
Simberloff, D. (1980). A succession of paradigms in ecology: essentialism to materialism and probabilism. Synthese, 43, 3–39.CrossRefGoogle Scholar
Sinclair, A. F. and Swain, D. P. (1996). Comment: spatial implications of a temperature-based growth model for Atlantic Cod (Gadus morhua) off the eastern coast of Canada. Canadian Journal of Fisheries and Aquatic Sciences, 53, 2909–2911.CrossRefGoogle Scholar
Sinclair, A. F., Swain, D. P. and Hanson, J. M. (2002a). Disentangling the effects of size-selective mortality, density, and temperature on length-at-age. Canadian Journal of Fisheries and Aquatic Sciences, 59, 372–382.CrossRefGoogle Scholar
Sinclair, M., Arnason, R., Csirke, J., et al. (2002b). Responsible fisheries in the marine ecosystem. Fisheries Research, 58, 255–265.CrossRefGoogle Scholar
Singleton, D. A. and Vo, L. K. (2002). Enantioselective synthesis without discrete optically active additives. Journal of the American Chemical Society, 124, 10 010–10 011.CrossRefGoogle ScholarPubMed
Siwoff, S., Hirdt, S., Hirdt, P. and Hirdt, T. (1990). The 1990 Elias Baseball Analyst. New York: Collier Books.Google Scholar
Skellam, J. G. (1951). Random dispersal in theoretical populations. Biometrika, 38, 196–218.CrossRefGoogle ScholarPubMed
Skorokhod, A. V., Hoppensteadt, F. C. and Salehi, H. (2002). Random Perturbation Methods. New York: Springer Verlag.CrossRefGoogle Scholar
Slepoy, A., Singh, R. R. P., Pázmándi, F., Kulkarni, R. V. and Cox, D. L. (2001). Statistical mechanics of prion diseases. Physical Review Letters, 87, 058101.CrossRefGoogle ScholarPubMed
Smith, G. C. and Wilkinson, D. (2002). Modelling disease spread in a novel host: rabies in the European badger Meles meles. Journal of Applied Ecology, 39, 865–874.CrossRefGoogle Scholar
Smith, M. C., Reeder, R. H. and Thomas, M. B. (1997). A model to determine the potential for biological control of Rottboellia cochinchinensis with the head smut Sporisorium ophiuri. Journal of Applied Ecology, 34, 388–398.CrossRefGoogle Scholar
Smith, R. D. (2000). Can't You Hear Me Callin'. The Life of Bill Monroe. Boston: Little, Brown and Company.Google Scholar
Smith, T. D. (1994). Scaling Fisheries. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Snow, C. P. (1965). The Two Cultures. Cambridge: Cambridge University Press.Google Scholar
Soai, K., Niwa, S. and Hori, H. (1990). Asymmetric self-catalytic reaction. Self-production of chiral 1-(3-pyridyl)alkanols as chiral self-catalysts in the enantioselective addition of dialkylzinc reagents to pyridine-3-caraldehyde. Journal of the Chemical Society, Chemical Communications, 1990, 982–983.CrossRefGoogle Scholar
Soh, S., Gunderson, D. R. and Ito, D. H. (2000). The potential role of marine reserves in the management of shortraker rockfish (Sebastest borealis) and rougheye rockfish (S. aleutianus) in the Gulf of Alaska. Fishery Bulletin, 99, 168–179.Google Scholar
Soule, M. E. (editor) (1987). Viable Populations for Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Sounes, H. (2001). Down the Highway. The Life of Bob Dylan. New York: Grove Press.Google Scholar
Southwood, T. R. E. (1978). Ecological Methods. London: Chapman and Hall.Google Scholar
Souza, P. D. (1998). On hyperbolic discounting and uncertain hazard rates. Proceedings of the Royal Society of London, B265, 2015–2020.CrossRefGoogle Scholar
Spielman, A. and D'Antonio, M. (2001). Mosquito. New York: Hyperion.Google Scholar
Stachel, J. (editor) (1998). Einstein's Miraculous Year. Five Papers that Changed the Face of Physics. Princeton, NJ: Princeton University Press.Google Scholar
Starr, P., Annala, J. H. and Hilborn, R. (1998). Contested stock assessment: two case studies. Canadian Journal of Fisheries and Aquatic Sciences, 55, 529–537.CrossRefGoogle Scholar
Stearns, S. C. (editor) (1999). Evolution in Health and Disease. New York: Oxford University Press.Google Scholar
Stearns, S. C. (2000). Daniel Bernoulli (1738): evolution and economics under risk. Journal of Biosciences, 25, 221–228.CrossRefGoogle Scholar
Stenseth, N. C., Bjørnstad, O. N., Falck, W., et al. (1999). Dynamics of coastal cod populations: intra- and intercohort density dependence and stochastic processes. Proceedings of the Royal Society of London, A266, 1645–1654.CrossRefGoogle Scholar
Stephens, D. W. and Krebs, J. R. (1986). Foraging Theory. Princeton, NJ: Princeton University Press.Google Scholar
Stewart, I. (2000). The Lorenz attractor exists. Nature, 406, 948–949.CrossRefGoogle ScholarPubMed
Stockwell, C. A., Hendry, A. P., Berg, O. K., Quinn, T. P. and Kinnison, M. T. (2003). Contemporary evolution meets conservation biology. Trends in Ecology and Evolution, 18, 94–101.CrossRefGoogle Scholar
Stolz, J. and Schnell, J. (editors) (1991). Trout. Harrisburg, PA: Stackpole Books.Google Scholar
Strand, E., Huse, G. and Giske, J. (2002). Artificial evolution of life history and behavior. American Naturalist, 159, 624–644.CrossRefGoogle ScholarPubMed
Stratonovich, R. L. (1963). Topics in the Theory of Random Noise. New York: Gordon and Breach.Google Scholar
Stratoudakis, Y., Fryer, R. J. and Cook, R. M. (1998). Discarding practices for commercial gadoids in the North sea. Canadian Journal of Fisheries and Aquatic Sciences, 55, 1632–1644.CrossRefGoogle Scholar
Strogatz, S. H. (1994). Nonlinear Dynamics and Chaos. Cambridge, MA: Westview.Google Scholar
Stroock, D. W. (2003). Markov Processes from K. Ito's Perspective. Princeton, NJ: Princeton University Press.CrossRefGoogle Scholar
Strunk, W. J. and White, E. B. (1979). The Elements of Style, 3rd edn. New York: MacMillan.Google Scholar
Sumpter, D. J. T. and Martin, J. (2004). The dynamics of virus epidemics in Varroa-infested honey bee colonies. Journal of Animal Ecology, 73, 51–63.CrossRefGoogle Scholar
Suter, G. W. (1996). Abuse of hypothesis testing statistics in ecological risk assessment. Human and Ecological Risk Assessment, 2, 331–347.CrossRefGoogle Scholar
Swansburg, E., Chaput, G., Moore, D., Caissie, D. and El-Jabi, N. (2002). Size variability of juvenile Atlantic salmon: links to environmental conditions. Journal of Fish Biology, 61, 661–683.CrossRefGoogle Scholar
Sylvain, G., Baalen, M. and Jansen, V. A. A. (2002). The evolution of parasite virulence, superinfection, and host resistance. American Naturalist, 159, 658–669.Google Scholar
Tanaka, M. M., Kumm, J. and Feldman, M. W. (2002). Coevolution of pathogens and cultural practices: a new look at behavioral heterogeneity in epidemics. Theoretical Population Biology, 62, 111–119.CrossRefGoogle Scholar
Tang, S. and Wang, Y. (2002). A parameter estimation program for the error-in-variable model. Ecological Modelling, 156, 225–236.CrossRefGoogle Scholar
Taper, M. L. and Lele, S. R. (2004). The Nature of Scientific Evidence. Statistical, Philosophical and Empirical Considerations. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Taylor, A. D. (1988a). Large-scale spatial structure and population dynamics in arthropod predator–prey systems. Annals Zoologia Fennici, 25, 63–74.Google Scholar
Taylor, A. D. (1988b). Parasitoid competition and the dynamics of host–parasitoid models. American Naturalist, 132, 417–436.CrossRefGoogle Scholar
Taylor, A. D. (1993). Aggregation, competition, and host–parasitoid dynamics: stability conditions don't tell it all. American Naturalist, 141, 501–506.CrossRefGoogle Scholar
Taylor, L. R. (1984). Assessing and interpreting the spatial distributions of insect populations. Annual Review of Entomology, 29, 321–357.CrossRefGoogle Scholar
Taylor, L. R., Woiwood, I. P. and Perry, J. N. (1978). The density-dependence of spatial behaviour and the rarity of randomness. Journal of Animal Ecology, 47, 383–406.CrossRefGoogle Scholar
Taylor, L. R., Woiwood, I. P. and Perry, J. N. (1979). The negative binomial as a dynamic ecological model for aggregation, and the density dependence of k. Journal of Animal Ecology, 48, 289–304.CrossRefGoogle Scholar
Taylor, L. R., Woiwood, I. P. and Perry, J. N. (1980). Variance and the large scale spatial stability of aphids, moths and birds. Journal of Animal Ecology, 49, 831–854.CrossRefGoogle Scholar
Haar, D. (1998). Master of Modern Physics. The Scientific Contributions of H. A. Kramers. Princeton, NJ: Princeton University Press.Google Scholar
Thom, R. (1972/1975). Stabilité structurelle et morphogenese (Structural Stability and Morphogenesis). Reading, MA: Benjamin.Google Scholar
Thompson, G. G. (1992). A Bayesian approach to management advice when stock-recruitment parameters are uncertain. Fishery Bulletin, 90, 561–573.Google Scholar
Thompson, G. G. (1993). A proposal for a threshold stock size and maximum fishing mortality rate. In Smith, S. J., Hunt, J. J. and Rivard, D. (editors) Risk Evaluation and Biological Reference Points for Fisheries Management (Canadian Special Publication of Fisheries and Aquatic Sciences 120). Ottawa: National Research Council of Canada.Google Scholar
Toft, C. A. and Mangel, M. (1991). From individuals to ecosystems: the papers of Skellam, Lindeman and Hutchinson. Bulletin of Mathematical Biology, 53, 121–134.CrossRefGoogle Scholar
Trumble, R. J. (1998). Northeast Pacific flatfish management. Journal of Sea Research, 39, 167–181.CrossRefGoogle Scholar
Turchin, P. (1998). Quantitative Analysis of Movement. Measuring and Modeling Population Redistribution in Animals and Plans. Sunderland, MA: Sinauer Associates.Google Scholar
Turelli, M. A., Schemske, D. W. and Bierzychudek, P. (2001). Stable two-allele polymorphisms maintained by fluctuating fitnesses and seed banks: protecting the blues in Linanthus parryae. Evolution, 55, 1283–1298.CrossRefGoogle ScholarPubMed
Turing, A. M. (1952). The chemical basis of morphogenesis. Proceedings of the Royal Society of London, B237, 37–72.CrossRefGoogle Scholar
Tyutyunov, Y., Senina, I., Jost, C. and Arditi, R. (2002). Risk assessment of the harvested pike-perch population of the Azov Sea. Ecological Modelling, 149, 297–311.CrossRefGoogle Scholar
Uhlenbeck, G. E. and Ornstein, L. S. (1930). On the theory of Brownian motion. Physical Review, 36, 823–841.CrossRefGoogle Scholar
Ulrich, C., Gallic, B., Dunn, M. R. and Gascuel, D. (2002a). A multi-species multi-fleet bioeconomic simulation model for the English Channel artisinal fisheries. Fisheries Research, 58, 379–401.CrossRefGoogle Scholar
Ulrich, C. and Marchal, P. (2002). Sensitivity of some biological reference points to shifts in exploitation patterns and inputs uncertainty for three North Sea demersal stocks. Fisheries Research, 58, 153–169.CrossRefGoogle Scholar
Ulrich, C., Pascoe, D., Sparre, P. J., Wilde, J.-W. and Marchal, P. (2002b). Influence of trends in fishing power on bioeconomics in the North Sea flatfish fishery regulated by catches or by effort quotas. Canadian Journal of Fisheries and Aquatic Sciences, 59, 829–843.CrossRefGoogle Scholar
Valleron, A.-J., Boelle, P.-Y., Will, R. and Cesbron, J.-Y. (2001). Estimation of epidemic size and incubation time based on age characteristics of vCJD in the United Kingdom. Science, 294, 1726–1728.CrossRefGoogle ScholarPubMed
Alphen, J. J. M. and Visser, M. E. (1990). Superparasitism as an adaptive strategy. Annual Review of Entomology, 35, 59–79.CrossRefGoogle ScholarPubMed
Alphen, J. J. M., Dijken, M. J. and Waage, J. K. (1987). A functional approach to superparasitism: host discrimination needs not to be learnt. Netherlands Journal of Zoology, 37, 167–179.CrossRefGoogle Scholar
Alphen, J. J. M., Bernstein, C. and Driessen, G. (2003). Information acquisition and time allocation in insect parasitoids. Trends in Ecology and Evolution, 18, 81–87.CrossRefGoogle Scholar
van Baalen, M. (2000). The evolution of parasitoid egg load. In Hochberg, M. E. and Ives, A. R. (editors). Parasitoid Population Biology. Princeton, NJ: Princeton University Press.Google Scholar
Kampen, N. G. (1977). A soluble model for diffusion in a bistable potential. Journal of Statistical Physics, 17, 71–88.CrossRefGoogle Scholar
Kampen, N. G. (1981a). Ito versus Stratonovich. Journal of Statistical Physics, 24, 175–187.CrossRefGoogle Scholar
Kampen, N. G. (1981b). Stochastic Processes in Physics and Chemistry. Amsterdam: North-Holland Publishing Company.Google Scholar
Oostenbrugge, J. A. E., Densen, W. L. T. and Machiels, M. A. M. (2001). Risk aversion in allocating fishing effort in a highly uncertain coastal fishery for pelagic fish, Moluccas, Indonesia. Canadian Journal of Fisheries and Aquatic Sciences, 58, 1683–1691.CrossRefGoogle Scholar
Randen, E. J. and Roitberg, B. D. (1996). The effect of egg load on superparasitism by the snowberry fly. Entomologia Experimentalis et Applicata, 79, 241–245.CrossRefGoogle Scholar
Winkle, W., Jager, H. I., Railsback, S. F., et al. (1998). Individual-based model of sympatric populations of brown and rainbow trout for instream flow assessment: model description and calibration. Ecological Modelling, 110, 175–207.CrossRefGoogle Scholar
Vasseur, D. A. and Yodzis, P. (2004). The color of environmental noise. Ecology, 85, 1146–1152.CrossRefGoogle Scholar
Vaupel, J. W., Carey, J. R., Christensen, K., et al. (1998). Biodemographic trajectories of longevity. Science, 280, 855–859.CrossRefGoogle ScholarPubMed
Vestergaard, N. (1996). Discard behavior, highgrading and regulation: the case of the Greenland Shrimp Fishery. Marine Resource Economics, 11, 247–266.CrossRefGoogle Scholar
Vestergaard, N., Squires, D. and Kirkley, J. (2003). Measuring capacity and capacity utilization in fisheries: the case of the Danish Gill-net fleet. Fisheries Research, 60, 357–368.CrossRefGoogle Scholar
Visser, M. E. and Rosenheim, J. A. (1998). The influence of competition between foragers on clutch size decisions in insect parasitoids. Biological Control, 11, 169–174.CrossRefGoogle Scholar
Bertalanffy, L. (1957). Quantitative laws in metabolism and growth. Quarterly Review of Biology, 32, 217–231.CrossRefGoogle Scholar
Voronka, R. and Keller, J. (1975). Asymptotic analysis of stochastic models in population genetics. Mathematical Biosciences, 25, 331–362.CrossRefGoogle Scholar
Vos, M. and Hemerik, L. (2003). Linking foraging behavior to lifetime reproductive success for an insect parasitoid: adaptation to host distributions. Behavioral Ecology, 14, 236–245.CrossRefGoogle Scholar
Vos, M., Hemerik, L. and Vet, L. E. M. (1998). Patch exploitation by the parasitoids Cotesia rubecula and Cotesia glomerata in multi-patch environments with different host distributions. Journal of Animal Ecology, 67, 774–783.CrossRefGoogle Scholar
Wachter, K. W. (1999). Evolutionary demographic models for mortality plateaus. Proceedings of the National Academy of Sciences, 96, 10 544–10 547.CrossRefGoogle ScholarPubMed
Wachter, K. W. and Finch, C. E. (editors) (1997). Between Zeus and the Salmon. The Biodemography of Longevity. Washington, DC: National Academy Press.Google Scholar
Wade, P. R. (2000). Bayesian methods in conservation biology. Conservation Biology, 14, 1308–1316.CrossRefGoogle Scholar
Waeckers, F. L. (1994). The effect of food deprivation on the innate visual and olfactory preferences in the parasitoid Cotesia rubecula. Journal of Insect Physiology, 40, 641–649.CrossRefGoogle Scholar
Wajnberg, E., Fauveergue, X. and Pons, O. (2000). Patch leaving decision rules and the marginal value theorem: an experimental analysis and a simulation model. Behavioral Ecology, 11, 577–586.CrossRefGoogle Scholar
Walde, S. J. and Murdoch, W. W. (1988). Spatial density dependence in parasitoids. Annual Review of Entomology, 33, 441–466.CrossRefGoogle Scholar
Walter, G. H. (1988). Activity patterns and egg production in Coccophagus bartletti, an aphelinid parasitoid of scale insects. Ecological Entomology, 13, 95–105.CrossRefGoogle Scholar
Walters, C. J. and Martell, S. J. D. (2004). Fisheries Ecology and Management. Princeton, NJ: Princeton University Press.Google Scholar
Waser, P. M., Elliott, L. F., Creel, N. M. and Creel, S. R. (1995). Habitat variation and viverrid demography. In Sinclair, A. R. E. and Arcese, P. (editors) Serengeti II: Dynamics, Management, and Conservation of an Ecosystem. Chicago, IL: University of Chicago Press, pp. 421–447.Google Scholar
Washburn, A. R. (1981). Search and Detection. Arlington, VA: Military Applications Section, Operations Research Society of America.Google Scholar
Watson, R. (1993). The Trout. A Fisherman's Natural History. Shrewsbury, UK: Swan Hill Press.Google Scholar
Watson, R. (1999). Salmon, Trout and Charr of the World. A Fisherman's Natural History. Shrewsbury, UK: Swan Hill Press.Google Scholar
Wegner, K., Kalbe, M., Kurtz, J., Reusch, T. and Milinski, M. (2003). Parasite selection for immunogenetic optimality. Science, 301, 1343.CrossRefGoogle ScholarPubMed
Weitz, J. S. and Fraser, H. B. (2001). Explaining mortality plateaus. Proceedings of the National Academy of Sciences, 98, 15 383–15 386.CrossRefGoogle Scholar
West, G. B., Brown, J. H. and Enquist, B. J. (1997). A general model for the origin of allometric scaling laws in biology. Science, 276, 122–126.CrossRefGoogle ScholarPubMed
West, G. B., Brown, J. H. and Enquist, B. J. (2001). A general model for ontogenetic growth. Nature, 413, 628–631.CrossRefGoogle ScholarPubMed
West, M. and Harrison, J. (1997). Bayesian Forecasting and Dynamic Models. New York: Springer Verlag.Google Scholar
West-Eberhard, M. J. (2003). Developmental Plasticity and Evolution. New York: Oxford University Press.Google Scholar
Wheatley, D. N. and Agutter, P. S. (1996). Historical aspects of the origin of diffusion theory in 19th-century mechanistic materialism. Perspectives in Biology and Medicine, 40, 139–155.CrossRefGoogle Scholar
White, A., Bowers, R. G. and Begon, M. (1996). Red/blue chaotic power spectra. Nature, 381, 198.CrossRefGoogle ScholarPubMed
White, P. C. L. and Harris, S. (1995). Bovine tuberculosis in badger (Meles meles) populations in southwest England: the use of a spatial stochastic simulation model to understand the dynamics of the disease. Philosophical Transactions of the Royal Society of London, B349, 391–413.CrossRefGoogle ScholarPubMed
Whittle, P. (1955). The outcome of a stochastic epidemic. A note on Bailey's paper. Biometrika, 42, 116–122.Google Scholar
Whittle, P. (1983). Optimization over Time. Dynamic Programming and Stochastic Control. Volume II. New York: John Wiley and Sons.Google Scholar
Wilcox, C. and Elderd, B. (2003). The effect of density-dependent catastrophes on population persistence time. Journal of Applied Ecology, 40, 859–871.CrossRefGoogle Scholar
Wilcox, C. and Possingham, H. (2002). Do life history traits affect the accuracy of diffusion approximations for mean time to extinction? Ecological Applications, 12, 1163–1179.CrossRefGoogle Scholar
Wilkinson, D. and Sherratt, T. N. (2001). Horizontally acquired mutualisms, an unsolved problem in ecology? Oikos, 92, 377–384.CrossRefGoogle Scholar
Williams, E. H. (2002). The effects of unaccounted discards and misspecified natural mortality on harvest policies based on estimates of spawners per recruit. North American Journal of Fisheries Management, 22, 311–325.2.0.CO;2>CrossRefGoogle Scholar
Williams, G. C. and Nesse, R. M. (1991). The dawn of Darwinian medicine. Quarterly Review of Biology, 66, 1–22.CrossRefGoogle ScholarPubMed
Williams, J. G. (2003). Sardine fishing in the early 20th century. Science, 300, 2032–2033.CrossRefGoogle ScholarPubMed
Wilmott, P. (1998). Derivatives. The Theory and Practice of Financial Engineering. Chichester: John Wiley and Sons.Google Scholar
Wilson, W. G., Morris, W. F. and Bronstein, J. L. (2003). Coexistence of mutualists and exploiters on spatial landscapes. Ecological Monographs, 73, 397–413.CrossRefGoogle Scholar
Wong, E. (1971). Stochastic Processes in Information and Dynamical Systems. New York: McGraw-Hill.Google Scholar
Wong, E. and Zakai, M. (1965). On the relation between ordinary and stochastic differential equations. International Journal of Engineering Science, 3, 213–229.Google Scholar
Wooster, W. S. (editor) (1988). Fishery Science and Management. Objectives and Limitations. New York: Springer Verlag.CrossRefGoogle Scholar
Wouk, H. (1962). Youngblood Hawke. New York: Doubleday.Google Scholar
Wouk, H. (1994). The Glory. New York: Little, Brown.Google Scholar
Wynia, M. K. and Gostin, L. (2002). The bioterrorist threat and access to health care. Science, 296, 1613.CrossRefGoogle ScholarPubMed
Yates, F. (1951). The influence of Statistical Methods for Research Workers on the development of the science of statistics. Journal of the American Statistical Association, 46, 19–34.Google Scholar
Yildiz, F. H. and Schoolnik, G. K. (1999). Vibrio cholerae O1 El Tor: Identification of a gene cluster required for the rugose colony type, exopolysaccharide production, chlorine resistance, and biofilm formation. Proceedings of the National Academy of Sciences, 96, 4028–4033.CrossRefGoogle ScholarPubMed
Young, T. P. (1994). Natural die-offs of large mammals: implications for conservation. Conservation Biology, 8, 410–418.CrossRefGoogle Scholar
Zabel, R. W., Harvey, C. J., Katz, S. L., Good, T. P. and Levin, P. S. (2003). Ecologically sustainable yield. American Scientist, 91, 150–157.Google Scholar
Zinnser, W. (1989). Writing to Learn. New York: Harper and Row.Google Scholar
Zinsser, W. (2001). On Writing Well. New York: Harper Collins.Google Scholar
Zuk, M. and Stoehr, A. M. (2002). Immune defense and host life history. American Naturalist, 160, S9–S22.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.

  • References
  • Marc Mangel, University of California, Santa Cruz
  • Book: The Theoretical Biologist's Toolbox
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511819872.011
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.

  • References
  • Marc Mangel, University of California, Santa Cruz
  • Book: The Theoretical Biologist's Toolbox
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511819872.011
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.

  • References
  • Marc Mangel, University of California, Santa Cruz
  • Book: The Theoretical Biologist's Toolbox
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511819872.011
Available formats
×