Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-rvbq7 Total loading time: 0 Render date: 2024-07-13T13:17:09.087Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  06 July 2018

Ann E. Hajek
Affiliation:
Cornell University, New York
Jørgen Eilenberg
Affiliation:
University of Copenhagen
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Natural Enemies
An Introduction to Biological Control
, pp. 402 - 425
Publisher: Cambridge University Press
Print publication year: 2018

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

Agrawal, A. A. (2000). Mechanisms, ecological consequences and agricultural implications of tri-trophic interactions. Current Opinion in Plant Biology, 3, 329335.Google Scholar
Agrios, G. N. (2005). Plant Pathology, 5th edn. San Diego, CA: Academic Press.Google Scholar
Albajes, R., Gullino, M. L., van Lenteren, J. C., & Elad, Y. (eds.) (1999). Integrated Pest and Disease Management in Greenhouse Crops. Dordrecht, Netherlands: Kluwer Academic.CrossRefGoogle Scholar
Andow, D. A. (1991). Vegetational diversity and arthropod population response. Annual Review of Entomology, 36, 561586.CrossRefGoogle Scholar
Andrews, J. H. & Harris, R. F. (1986). r- and K-selection and microbial ecology. Advances in Microbial Ecology, 9, 99147.CrossRefGoogle Scholar
Annecke, D. P., Karny, M., & Burger, W. A. (1969). Improved biological control of the prickly pear, Opuntia megacantha Salm-Dyck, in South Africa through use of an insecticide. Phytophylactica, 1, 913.Google Scholar
Aquilino, K. M., Cardinale, B. J., & Ives, A. R. (2005). Reciprocal effects of host plant and natural enemy diversity on herbivore suppression: An empirical study of a model tritrophic system. Oikos, 108, 275282.CrossRefGoogle Scholar
Askew, R. R. (1971). Parasitic Insects. London: Heinemann Educational Books.Google Scholar
Askew, R. R. (1975). The organisation of chalcid-dominated parasitoid communities centered upon endophytic hosts. In Evolutionary Strategies of Parasitoids, ed. Price, P. W., pp. 130153. New York: Plenum Press.Google Scholar
Axtell, R. C. (1986). Fly Control in Confined Livestock and Poultry Production. Greensboro, NC: CIBA-GEIGY Corp.Google Scholar
Bajwa, W. I. & Kogan, M. (2002). Compendium of IPM definitions (CID) – What is IPM and how is it defined in the worldwide literature? IPPC Publication No. 998. Integrated Plant Protection Center (IPPC), Oregon State University, Corvallis, OR.Google Scholar
Baker, K. F. & Griffin, G. J. (1995). Molecular strategies for biological control of fungal plant pathogens. In Novel Approaches to Integrated Pest Management, ed. Reuveni, R., pp. 153182. Boca Raton, FL: CRC Press.Google Scholar
Balciunas, J. K. (2000). A proposed code of best practices for classical biological control of weeds. In Proceedings of the X International Symposium on Biological Control of Weeds, 5–9 July 1999, Bozeman, Montana, ed. Spencer, N. R., pp. 435436. Bozeman: Montana State University.Google Scholar
Banfield-Zanin, J. A. & Leather, S. R. (2016). Prey-mediated effects of drought on the consumption rates of coccinellid predators of Elatobium abietinum. Insects, 7 (4), 49.CrossRefGoogle ScholarPubMed
Barbosa, P. (ed.) (1998). Conservation Biological Control. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Barbosa, P., Gross, P., & Kemper, J. (1991). Influence of plant allelochemicals on the tobacco hornworm and its parasitoid, Cotesia congregata. Ecology, 72, 15671575.CrossRefGoogle Scholar
Barbosa, P., Saunders, J. A., Kemper, J., Trumbule, R., Olechno, J., & Martinat, P. (1986). Plant allelochemicals and insect parasitoids: Effects of nicotine on Cotesia congregata (Say) (Hymenoptera: Braconidae) and Hyposoter annulipes (Cresson) (Hymenoptera: Ichneumonidae). Journal of Chemical Ecology, 12, 13191328.Google Scholar
Barron, G. L. (1979). Observations on predatory fungi. Canadian Journal of Botany, 57, 187193.Google Scholar
Bateman, R. P. (2000). Rational pesticide use: An alternative escape from the treadmill? Biocontrol News & Information, 21, 96100.Google Scholar
Baum, J. A., Johnson, T. B., & Carlton, B. C. (1998). Bacillus thuringiensis, natural and recombinant bioinsecticide products. In Biopesticides, Use and Delivery, ed. Hall, F. R. & Menn, J. J., pp. 189209. Totowa, NJ: Humana Press.Google Scholar
Becker, N. (2000). Bacterial control of vector-mosquitoes and black flies. In Entomopathogenic Bacteria: From Laboratory to Field Application, ed. Charles, J.-F., Delécluse, A., & Nielsen-LeRoux, C., pp. 383398. Dordrecht, Netherlands: Kluwer Academic Publishers.CrossRefGoogle Scholar
Bedding, R. A. (1993). Biological control of Sirex noctilio using the nematode Deladenus siricidicola. In Nematodes and the Biological Control of Insect Pests, ed. Bedding, R., Akhurst, R., & Kaya, H., pp. 1120. East Melbourne, Australia: CSIRO.Google Scholar
Bellows, T. S. & Fisher, T. W. (eds.) (1999). Handbook of Biological Control. San Diego, CA: Academic Press.Google Scholar
Bellows, T. S., Paine, T. D., Gould, J. R., Bezark, L. G., & Ball, J. C. (1992). Biological control of ash whitefly: A success in progress. California Agriculture, 46 (1), 2428.Google Scholar
Benbrook, C. M. (1996). Pest Management at the Crossroads. Yonkers, NY: Consumers Union.Google Scholar
Berenbaum, M. R. (1995). Turnabout is fair play: Secondary roles for primary compounds. Journal of Chemical Ecology, 21, 925940.Google Scholar
Bergelson, J. & Crawley, M. J. (1989). The theory and practice of biological control. Comments on Modern Biology, C, Comments on Theoretical Biology, 1, 197215.Google Scholar
Berner, D. K., Bruckart, W. L., Cavin, C. A., Michael, J. L., Carter, M. L., & Luster, D. G. (2009). Best linear unbiased prediction of host-range of the facultative parasite Colletotrichum gloeosporioides f. sp. salsolae, a potential biological control agent of Russian thistle. Biological Control, 51, 158168.Google Scholar
Biddinger, D. & Butler, B. (2013). Biological control of mites in Pennsylvania and Maryland apple orchards. Fruit Times, July 26. Retrieved from http://extension.psu.edu/plants/tree-fruit/news/2013/biological-control-of-mites-in-pennsylvania-and-maryland-apple-orchards.Google Scholar
Blossey, B. & Hunt-Joshi, T. R. (2003). Belowground herbivory by insects: Influence on plants and aboveground herbivores. Annual Review of Entomology, 48, 521547.CrossRefGoogle ScholarPubMed
Blossey, B., Skinner, L. C., & Taylor, J. (2001). Impact and management of purple loosestrife (Lythrum salicaria) in North America. Biodiversity and Conservation, 10, 17871807.Google Scholar
Boettner, G. H., Elkinton, J. S., & Boettner, C. J. (2000). Effects of a biological control introduction on three nontarget native species of saturniid moths. Conservation Biology, 14, 17981806.Google Scholar
Borriss, R., Chen, X.-H., Rueckert, C., Blom, J., Becker, A., Baumgarth, B., Fan, B., Pukall, R., Schumann, P., Spröer, C., et al. (2011). Relationship of Bacillus amyloliquefaciens clades associated with strains DSM 7T and FZB42T: A proposal for Bacillus amyloliquefaciens subsp. amyloliquefaciens subsp. nov. and Bacillus amyloliquefaciens subsp. plantarum subsp. nov. based on complete genome sequence comparisons. International Journal of Systematic and Evolutionary Microbiology, 61, 17861801.CrossRefGoogle ScholarPubMed
Boucias, D. G. & Pendland, J. C. (1998). Principles of Insect Pathology. Boston, MA: Kluwer Academic Publishers.CrossRefGoogle Scholar
Bourchier, R. S. & Van Hezewijk, B. H. (2013). Euphorbia esula L., Leafy spurge (Euphorbiaceae). In Biological Control Programmes in Canada 2001–2012, ed. Mason, P. G. & Gillespie, D. R.., pp. 315320. Wallingford, UK: CABI Publishing.Google Scholar
Brennan, E. B. (2013). Agronomic aspects of strip intercropping lettuce with alyssum for biological control of aphids. Biological Control, 65, 302311.Google Scholar
Briese, D. T. & McLaren, D. A. (1997). Community involvement in the distribution and evaluation of biological control agents: Landcare and similar groups in Australia. Biocontrol News & Information, 18, 39N-49N.Google Scholar
Brown, G. C. (1987). Modeling. In Epizootiology of Insect Diseases, ed. Fuxa, J. R. & Tanada, Y., pp. 4368. New York: John Wiley & Sons.Google Scholar
Bugg, R. L. & Pickett, C. H. (1998). Introduction: Enhancing biological control – habitat management to promote natural enemies of agricultural pests. In Enhancing Biological Control: Habitat Management to Promote Natural Enemies of Agricultural Pests, ed. Pickett, C. H. & Bugg, R. L., pp. 123. Berkeley, CA: University of California Press.Google Scholar
Bull, D. L. & Menn, J. J. (1990). Strategies for managing resistance to insecticides in Heliothis pests of cotton. In Managing Resistance to Agrochemicals, ed. Green, M. B., LeBaron, H. M., & Moberg, W. K., pp. 118133. Washington, DC: American Chemical Society.CrossRefGoogle Scholar
Burdon, J. J. & Marshall, D. R. (1981). Biological control and the reproductive mode of weeds. Journal of Applied Ecology, 18, 649658.CrossRefGoogle Scholar
Burges, H. D. (ed.) (1998). Formulation of Microbial Biopesticides. Dordrecht, Netherlands: Kluwer Academic Publishers.CrossRefGoogle Scholar
Butler, L., Zivkovich, C., & Sample, B. E. (1995). Richness and abundance of arthropods in the oak canopy of West Virginia's Eastern Ridge and Valley Section during a study of impact of Bacillus thuringiensis with emphasis on macrolepidoptera larvae. Bulletin of the Agricultural Forest Experiment Station, West Virginia University, 711, 119.Google Scholar
Cade, W. (1975). Acoustically orienting parasitoids: Fly phonotaxis to cricket song. Science, 190, 13121313.Google Scholar
Caltagirone, L. E. (1981). Landmark examples in classical biological control. Annual Review of Entomology, 26, 213232.CrossRefGoogle Scholar
Caltagirone, L. E., Shea, K. P., & Finney, G. L. (1964). Parasites to aid control of navel orangeworm. California Agriculture, 18(1), 1012.Google Scholar
Campbell, R. (1989). Biological Control of Microbial Plant Pathogens. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Cartwright, B. & Kok, L. T. (1990). Feeding by Cassida rubiginosa (Coleoptera: Chrysomelidae) and the effects of defoliation on growth of musk thistles. Journal of Entomological Science, 25, 538547.CrossRefGoogle Scholar
Center, T. D., Hill, M. P., Cordo, H., & Julien, M. H. (2002). Waterhyacinth. In Biological Control of Invasive Plants in the Eastern United States, ed. Van Driesche, R., Blossey, B., Hoddle, M., Lyon, S., & Reardon, R., pp. 4164. Morgantown, WV: USDA Forest Service Publication, FHTET-2002-04.Google Scholar
Chandler, J. M. (1980). Assessing losses caused by weeds. In Proceedings of the E. C. Stakman Communication Symposium, ed. Aldrich, R. J.. University of Minnesota Agricultural Experiment Station, Miscellaneous Publication 7.Google Scholar
Chaney, W. E. (1998). Biological control of aphids in lettuce using in-field insectaries. In Enhancing Biological Control: Habitat Management to Promote Natural Enemies of Agricultural Pests, ed. Pickett, C. H. & Bugg, R. L., pp. 7383. Berkeley: University of California Press.Google Scholar
Charudattan, R. (2005). Ecological, practical, and political inputs into selection of weed targets: What makes a good biological control target? Biological Control, 35, 183196.CrossRefGoogle Scholar
Charudattan, R. (2010). A reflection on my research in weed biological control: Using what we have learned for future applications. Weed Technology, 24, 208217.CrossRefGoogle Scholar
Charudattan, R. (2015). Weed control with microbial bioherbicides. In Weed Science for Sustainable Agriculture, Environment and Biodiversity, ed. Rao, A. N. & Yaduraju, N. T., Vol. 1, pp. 7996. Proceedings of the Plenary and Lead Papers of the 25th Asian-Pacific Weed Science Society Conference, Hyderabad, India.Google Scholar
Clarke, A. R. & Walter, G. H. (1995). “Strains” and the classical biological control of insect pests. Canadian Journal of Zoology, 73, 17771790.Google Scholar
Cloyd, R. A. (2012). Indirect effects of pesticides on natural enemies. In Pesticides – Advances in Chemical and Botanical Pesticides, ed. Soundararajan, R. P., pp. 127150. Rijeka, Croatia: InTech.Google Scholar
Cock, M. J. W., Murphy, S. T., Kairo, M. T. K., Thompson, E., Murphy, R. J., & Francis, A. W. (2016). Trends in the classical biological control of insect pests by insects: An update of the BIOCAT database. BioControl, 61, 349363.Google Scholar
Cock, M. J. W., van Lenteren, J. C., Brodeur, J., Barratt, B. I. P., Bigler, F., Bolckmans, K., Cônsoli, F. L., Haas, F., Mason, P. G., & Parra, J. R. P. (2010). Do new Access and Benefit Sharing procedures under the Convention on Biological Diversity threaten the future of biological control? BioControl, 55, 199218.Google Scholar
Coll, M. (1998). Parasitoid activity and plant species composition in intercropped systems. In Enhancing Biological Control: Habitat Management to Promote Natural Enemies of Agricultural Pests, ed. Pickett, C. H. & Bugg, R. L., pp. 85120. Berkeley: University of California Press.Google Scholar
Cook, R. J. (1993). Making greater use of introduced microorganisms for biological control of plant pathogens. Annual Review of Phytopathology, 31, 5380.Google Scholar
Coppel, H. C. & Mertins, J. W. (1977). Biological Insect Pest Suppression. Berlin: Springer-Verlag.Google Scholar
Costa, A. S. & Müller, G. W. (1980). Tristeza control by cross protection. Plant Disease, 64, 538541.Google Scholar
Crawley, M. J. (1989). The successes and failures of weed biocontrol using insects. Biocontrol News & Information, 10, 213223.Google Scholar
Cruz, Y. P. (1981). A sterile defender morph in a polyembryonic hymenopterous parasite. Nature, 294, 446447.Google Scholar
Culliney, T. W. (2005). Benefits of classical biological control for managing invasive plants. Critical Reviews in Plant Sciences, 24, 131150.Google Scholar
Dahlsten, D. L. & Garcia, R. (eds.) (1989). Eradication of Exotic Pests. New Haven, CT: Yale University Press.Google Scholar
Darwin, C. (1859). On the Origin of Species. London: John Murray & Sons.Google Scholar
Davies, K. G., Flynn, C. A., Laird, V., & Kerry, B. R. (1990). The life-cycle, population dynamics and host specificity of a parasite of Heterodera avenae, similar to Pasteuria penetrans. Revue de Nématologie, 13, 303309.Google Scholar
Deacon, J. W. (1991). Significance of ecology in the development of biocontrol agents against soil-borne plant pathogens. Biocontrol Science and Technology, 1, 520.Google Scholar
DeBach, P. H. (1964). Biological Control of Insect Pests and Weeds. New York: Reinhold Publishing Corporation.Google Scholar
DeBach, P. (1974). Biological Control by Natural Enemies. London: Cambridge University Press.Google Scholar
DeBach, P. & Rosen, D. (1991). Biological Control by Natural Enemies. Cambridge: Cambridge University Press.Google Scholar
DeBach, P., Rosen, D., & Kennett, C. E. (1971). Biological control of coccids by introduced natural enemies. In Biological Control, ed. Huffaker, C. B., pp. 165194. New York: Plenum Press.Google Scholar
DeBach, P. & Sundby, R. A. (1963). Competitive displacement between ecological homologues. Hilgardia, 34, 105166.CrossRefGoogle Scholar
De Barro, P. J. & Coombs, M. T. (2009). Post-release evaluation of Eretmocerus hayati Zolnerowich and Rose in Australia. Bulletin of Entomological Research, 99, 193206.Google Scholar
De Barro, P. J., Liu, S.-S., Boykin, L. M., & Dinsdale, A. B. (2011). Bemisia tabaci: A statement of species status. Annual Review of Entomology, 56, 119.Google Scholar
Delfosse, E. S. & Cullen, J. M. (1980). New activities in biological control of weeds in Australia. II. Echium plantagineum: Curse or salvation? In Proceedings of the Fifth International Symposium on Biological Control of Weeds, Brisbane, Australia, pp. 563574.Google Scholar
De Moraes, C. M., Lewis, W. J., Paré, P. W., Alborn, H. T., & Tumlinson, J. H. (1998). Herbivore-infested plants selectively attract parasitoids. Science, 393, 570573.Google Scholar
Dennill, G. B. (1985). The effect of the gall wasp Trichilogaster acaciaelongifoliae (Hymenoptera: Pteromalidae) on reproductive potential and vegetative growth of the weed Acacia longifolia. Agriculture, Ecosystems & Environment, 14, 5361.CrossRefGoogle Scholar
Denoth, M., Frid, L., & Myers, J. H. (2002). Multiple agents in biological control: improving the odds? Biological Control, 24, 2030.Google Scholar
de Oliviera, M. R. V. (1998). South America. In Insect Viruses and Pest Management, ed. Hunter-Fujita, F. R., Entwistle, P. F., Evans, H. F., & Crook, N. E., pp. 339355. Chichester, UK: John Wiley & Sons.Google Scholar
Dietrick, E. J., Schlinger, E. I., & Garber, M. J. (1960). Vacuum cleaner principle applied in sampling insect populations in alfalfa fields by new machine method. California Agriculture, 14 (1), 911.Google Scholar
Di Giallonardo, F. & Holmes, E. C. (2015). Exploring host-pathogen interactions through biological control. PLOS Pathogens, 11(6), e1004865.Google Scholar
DiTommaso, A., Averill, K. M., Hoffmann, M. P., Fuchsberg, J. R., & Losey, J. E. (2016). Integrating insect, resistance, and floral resource management in weed control decision-making. Weed Science, 64, 743756.Google Scholar
Dixon, A. F. G. (2000). Insect Predator–Prey Dynamics: Ladybird Beetles and Biological Control. Cambridge: Cambridge University Press.Google Scholar
Dixon, B. (1994). Keeping an eye on B. thuringiensis. BioTechnology, 12, 435.Google Scholar
Dobson, A. P. (1988). Restoring island ecosystems: The potential of parasites to control introduced mammals. Conservation Biology, 2, 3139.CrossRefGoogle Scholar
Douglas, A. E. (1998). Nutritional interactions in insect-microbial symbioses: Aphids and their symbiotic bacteria Buchnera. Annual Review of Entomology, 43, 1737.Google Scholar
Duetting, P. S. (2002). Effect of field pea surface wax variation on infection of the pea aphid by the fungal pathogen, Pandora neoaphidis. MS thesis, University of Idaho.Google Scholar
Dyer, L. A. (1995). Tasty generalists and nasty specialists? Antipredator mechanisms in tropical lepidopteran larvae. Ecology, 76, 14831496.Google Scholar
Dyer, L. A. & Gentry, G. (1999). Predicting natural-enemy responses to herbivores in natural and managed systems. Ecological Applications, 9, 402408.Google Scholar
Ehler, L. E. (1998). Conservation biological control: Past, present, and future. In Conservation Biological Control, ed. Barbosa, P., pp. 18. San Diego, CA: Academic Press.Google Scholar
Eigenbrode, S. D., Castagnola, T., Roux, M.-B., & Steljes, L. (1996). Mobility of three generalist predators is greater on cabbage with glossy leaf wax than on cabbage with a wax bloom. Entomologia Experimentalis et Applicata, 81, 335343.CrossRefGoogle Scholar
Eilenberg, J., Enkegaard, A., Vestergaard, S., & Jensen, B. (2000). Biocontrol of pests on plant crops in Denmark: Present status and future potential. Biocontrol Science and Technology, 10, 703716.Google Scholar
Eilenberg, J., Hajek, A., & Lomer, C. (2001). Suggestions for unifying the terminology in biological control. BioControl, 46, 387400.Google Scholar
Elad, Y., Chet, I., Boyle, P., & Henis, Y. (1983). Parasitism of Trichoderma spp. on Rhizoctonia solani and Sclerotium rolfsii – scanning electron microscopy and fluorescence microscopy. Phytopathology, 73, 8588.Google Scholar
Elkinton, J. S., Healy, W. M., Buonaccorsi, J. P., Boettner, G. H., Hazzard, A. M., & Smith, H. R. (1996). Interactions among gypsy moths, white-footed mice, and acorns. Ecology, 77, 23322342.Google Scholar
Embree, D. G. (1966). The role of introduced parasites in the control of the winter moth in Nova Scotia. Canadian Entomologist, 98, 11591168.Google Scholar
Engelstädter, J. & Hurst, G. D. D. (2009).The ecology and evolution of microbes that manipulate host reproduction. Annual Review of Ecology, Evolution and Systematics, 40, 127149.Google Scholar
Eriksson, J., Khortstam, K., & Ryvarden, L. (1981). The Corticiaceae of North Europe. Oslo, Norway: Fungiflora.Google Scholar
Essig, E. O. (1942). College Entomology. New York: Macmillan Co.Google Scholar
Eubanks, M. D., Blackwell, S. A., Parish, C. J., DeLamar, Z. D., & Hull-Sanders, H. (2002). Intraguild predation of beneficial arthropods by red imported fire ants in cotton. Environmental Entomology, 31, 11681174.Google Scholar
Evans, E. W. & England, S. (1996). Indirect interactions in biological control of insects: Pests and natural enemies in alfalfa. Ecological Applications, 6, 920930.Google Scholar
Federici, B. (1999). Bacillus thuringiensis in biological control. In Handbook of Biological Control, ed. Bellows, T. S. & Fisher, T. W., pp. 575593. San Diego, CA: Academic Press.Google Scholar
Feitelson, J. S., Payne, J., & Kim, L. (1992). Bacillus thuringiensis: Insects and beyond. Bio/Technology, 10, 271275.Google Scholar
Felske, A. & Akkermans, A. D. L. (1998). Spatial homogeneity of abundant bacterial 16S rRNA molecules in grassland soils. Microbial Ecology, 36, 3136.Google Scholar
Fenner, F. & Fantini, B. (1999). Biological Control of Vertebrate Pests: The History of Myxomatosis – An Experiment in Evolution. Wallingford, UK: CABI Publishing.Google Scholar
Fenner, F. & Myers, K. (1978). Myxoma virus and myxomatosis in retrospect: The first quarter century of a new disease. In Viruses and Environment, ed. Kurstak, E. & Maramorosch, K., pp. 539570. New York: Academic Press.Google Scholar
Fields, P. G. & White, N. D. G. (2002). Alternatives to methyl bromide treatments for stored-product and quarantine insects. Annual Review of Entomology, 47, 331359.Google Scholar
Finch, S. & Collier, R. H. (2000). Host-plant selection by insects – a theory based on “appropriate/inappropriate landings” by pest insects of cruciferous plants. Entomologia Experimentalis et Applicata, 96, 91102.Google Scholar
Fischer, M. J., Havill, N. P., Brewster, C. C., Davis, G. A., Salom, S. M., & Kok, L. T. (2015). Field assessment of hybridization between Laricobius nigrinus and L. rubidus, predators of Adelgidae. Biological Control, 82, 16.Google Scholar
Flint, M. L. & Dreistadt, S. H. (1998). Natural Enemies Handbook: The Illustrated Guide to Biological Control. Berkeley: University of California Statewide IPM Project Publication 3386.Google Scholar
Flint, M. L. & Gouveia, P. (2001). IPM in Practice: Principles and Methods of Integrated Pest Management. Oakland: University of California, Division of Agriculture and Natural Resources.Google Scholar
Flint, M. L. & van den Bosch, R. (1981). Introduction to Integrated Pest Management. New York: Plenum Press.Google Scholar
Follett, P. A., Duan, J., Messing, R. H., & Jones, V. P. (2000). Parasitoid drift after biological control introductions: Re-examining Pandora's box. American Entomologist, 46, 8294.Google Scholar
Food and Agriculture Organization (UN). (2018). AGP – Integrated Pest Management. Retrieved from http://www.fao.org/agriculture/crops/thematic-sitemap/theme/pests/ipm/en/.Google Scholar
Fowler, S. V., Syrett, P., & Hill, R. L. (2000). Success and safety in the biological control of environmental weeds in New Zealand. Austral Ecology, 25, 553562.Google Scholar
Friedman, M. J. (1990). Commercial production and development. In Entomopathogenic Nematodes for Biological Control, ed. Gaugler, R. & Kaya, H. K., pp. 153172. Boca Raton, FL: CRC Press.Google Scholar
Frost, R. (1936). A Further Range. New York: Henry Hold & Co.Google Scholar
Fuester, R., Hajek, A. E., Schaefer, P., & Elkinton, J. S. (2014). Biological control of Lymantria dispar. In The Use of Classical Biological Control to Preserve Forests in North America, ed. Van Driesche, R. G. & Reardon, R., pp. 4982. Morgantown, WV: USDA Forest Service, Forest Health Technology Enterprise Team, FHTET-2013-02.Google Scholar
Fulbright, D. W. (1999). Hypovirulence to control fungal pathogenesis. In Handbook of Biological Control, ed. Bellows, T. S. & Fisher, T. W., pp. 691712. San Diego, CA: Academic Press.Google Scholar
Funasaki, G. Y., Lai, P.-Y., Nakahara, L. M., Beardsley, J. W., & Ota, A. K. (1988). A review of biological control introductions in Hawaii: 1890 to 1985. Proceedings of the Hawaiian Entomological Society, 28, 105160.Google Scholar
Fuxa, J. R. (1987). Ecological considerations for the use of entomopathogens in IPM. Annual Review of Entomology, 32, 225251.Google Scholar
Fuxa, J. R. (1998). Environmental manipulation for microbial control of insects. In Conservation Biological Control, ed. Barbosa, P., pp. 255268. San Diego, CA: Academic Press.Google Scholar
Gage, S. H. & Haynes, D. L. (1975). Emergence under natural and manipulated conditions of Tetrastichus julis, an introduced larval parasite of the cereal leaf beetle, with reference to regional population management. Environmental Entomology, 4, 425434.Google Scholar
Garcia, R., Caltagirone, L. E., & Gutierrez, A. P. (1988). Comments on a redefinition of biological control. BioScience, 38, 692694.Google Scholar
Garcia, R. & Legner, E. F. (1999). Biological control of medical and veterinary pests. In Handbook of Biological Control, ed. Bellows, T. S. & Fisher, T. W., pp. 935953. San Diego, CA: Academic Press.Google Scholar
Gardiner, M. M., Landis, D. A. Gratton, C., DiFonzo, C. D., O'Neal, M., Chacon, J. M., et al. (2009). Landscape diversity enhances biological control of an introduced crop pest in the north-central USA. Ecological Applications, 19, 143154.Google Scholar
Garnas, J. R., Hurley, B. P., Slippers, B., & Wingfield, M. J. (2012). Biological control of forest plantation pests in an interconnected world requires greater international focus. International Journal of Pest Management, 58, 211223.Google Scholar
Gause, G. F. (1934). The Struggle for Existence. Baltimore, MD: Williams and Wilkins.Google Scholar
Gelernter, W. D. & Lomer, C. J. (2000). Success in biological control of above-ground insects by pathogens. In Biological Control: Measures of Success, ed. Gurr, G. & Wratten, S., pp. 297322. Dordrecht, Netherlands: Kluwer Academic Publishers.Google Scholar
Gettys, L. A., Tipping, P. W., Della Torre, C. J., III, Sardes, S. N., & Thayer, K. M. (2014). Can herbicide usage be reduced by practicing IPM for waterhyacinth (Eichhornia crassipes) control? Proceedings of the Florida State Horticultural Society, 127, 14.Google Scholar
Glare, T., Caradus, J., Gelernter, W., Jackson, T., Keyhani, N., Köhl, J., Marrone, P., Morin, L., & Stewart, A. (2012). Have biopesticides come of age? Trends in Biotechnology, 30, 250258.Google Scholar
Glare, T. R. & O'Callaghan, M. (2000). Bacillus thuringiensis: Biology, Ecology and Safety. Chichester, UK: Wiley & Sons.Google Scholar
Godwin, P. A. & ODell, T. M. (1981). Intensive laboratory and field evaluations of individual species: Blepharipa pratensis (Meigen) (Diptera: Tachinidae). In The Gypsy Moth: Research Toward Integrated Pest Management, ed. Doane, C. C. & McManus, M. L., pp. 375394. Washington, DC: USDA, Forest Service, Technical Bulletin 1584.Google Scholar
Goeden, R. D. & Andrés, L. A. (1999). Biological control of weeds in terrestrial and aquatic environments. In Handbook of Biological Control, ed. Bellows, T. S. & Fisher, T. W., pp. 871890. San Diego, CA: Academic Press.Google Scholar
Goldberg, L. J. & Margalit, J. (1977). A bacterial spore demonstrating rapid larvicidal activity against Anopheles sergentii, Uranotaenia unguiculata, Culex univitattus, Aedes aegypti, and Culex pipiens. Mosquito News, 37, 355358.Google Scholar
Gordh, G., Legner, E. F., & Caltagirone, L. E. (1999). Biology of parasitic Hymenoptera. In Handbook of Biological Control, ed. Bellows, T. S. & Fisher, T. W., pp. 355381. San Diego, CA: Academic Press.Google Scholar
Gould, J., Hoelmer, K., & Goolsby, J. (eds.) (2008). Classical Biological Control of Bemisia tabaci in the United States – A Review of Interagency Research and Implementation. Dordrecht, Netherlands: Springer.Google Scholar
Graham, F. (1970). Since Silent Spring. Boston, MA: Houghton-Mifflin.Google Scholar
Gray, M. E., Ratcliffe, S. T., & Rice, M. E. (2009). The IPM paradigm: Concepts, strategies and tactics. In Integrated Pest Management: Concepts, Tactics, Strategies and Case Studies, ed. Radcliffe, E. B., Hutchison, W. E. & Cancelado, R. E., pp. 113. Cambridge: Cambridge University Press.Google Scholar
Grbić, M., Ode, P. J., & Strand, M. R. (1992). Sibling rivalry and brood sex ratios in polyembryonic wasps. Nature, 360, 254255.Google Scholar
Greathead, D. J. (1994). History of biological control. Antenna, 18, 187199.Google Scholar
Grewal, P. S. (2002). Formulation and application technology. In Entomopathogenic Nematology, ed. Gaugler, R., pp. 265287. Wallingford, UK: CABI Publishing.Google Scholar
Grime, J. P. (1977). Evidence for the existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. American Naturalist, 111, 11691194.CrossRefGoogle Scholar
Griswold, G. H. (1929). On the bionomics of a primary parasite and of two hyperparasites of the geranium aphid. Annals of the Entomological Society of America, 22, 438457.Google Scholar
Gross, P. (1991). Influence of target pest feeding niche on success rates in classical biological control. Environmental Entomology, 20, 12171227.Google Scholar
Gullan, P. J. & Cranston, P. S. (2000). The Insects: An Outline of Entomology, 2nd edn. Oxford: Blackwell Science.Google Scholar
Gurr, G. M. & Wratten, S. D. (eds.) (2000). Biological Control: Measures of Success. Dordrecht, Netherlands: Kluwer Academic Publishers.Google Scholar
Gurr, G. M., Wratten, S. D., Landis, D. A., & You, M. (2017). Habitat management to suppress pest populations: Progress and prospects. Annual Review of Entomology, 62, 91109.Google Scholar
Gwynn, R. L. (ed.) (2014). A World Compendium – The Manual of Biocontrol Agents, 5th edn. Alton, UK: British Crop Protection Council.Google Scholar
Hagen, K. S., Sawall, E. F., Jr., & Tassen, R. L. (1970). The use of food sprays to increase effectiveness of entomophagous insects. Proceedings of the Tall Timbers Conference on Ecological Animal Control by Habitat Management, 2, 5981.Google Scholar
Hajek, A. E. (1997). Fungal and viral epizootics in gypsy moth (Lepidoptera: Lymantriidae) populations in central New York. Biological Control, 10, 5868.Google Scholar
Hajek, A. E. (1999). Pathology and epizootiology of the lepidoptera-specific mycopathogen Entomophaga maimaiga. Microbiology and Molecular Biology Reviews, 63, 814835.Google Scholar
Hajek, A. E. (2004). Natural Enemies: An Introduction to Biological Control. Cambridge: Cambridge University Press.Google Scholar
Hajek, A. E., Delalibera, I., Jr., & Butler, L. (2003). Entomopathogenic fungi as classical biological control agents. In Environmental Impacts of Microbial Insecticides, ed. Hokkanen, H. M. T. & Hajek, A. E., pp. 1534. Dordrecht, Netherlands: Kluwer Academic Publishers.Google Scholar
Hajek, A. E., Gardescu, S., & Delalibera, I., Jr. (2016). Classical Biological Control of Insects and Mites: A Worldwide Catalogue of Pathogen and Nematode Introductions. USDA Forest Service, FHTET-2016-06.Google Scholar
Hajek, A. E., Hurley, B. P., Kenis, M., Garnas, J. R., Bush, S. J., Wingfield, M. J., van Lenteren, J. C., & Cock, M. J. W. (2016). Exotic biological control agents: a solution or contribution to arthropod invasions? Biological Invasions, 18, 953969.Google Scholar
Hajek, A. E., McManus, M. L., & Delalibera, I, Jr. (2007). A review of introductions of pathogens and nematodes for classical biological control of insects and mites. Biological Control, 41, 113.Google Scholar
Hajek, A. E., Wraight, S. P., & Vandenberg, J. D. (2001). Control of arthropods using pathogenic fungi. In Bio-Exploitation of Fungi, ed. Pointing, S. B. & Hyde, K. D., pp. 309347. Hong Kong: Fungal Diversity Press.Google Scholar
Hall, F. R. & Menn, J. J. (1999). Biopesticides, Use and Delivery. Totowa, NJ: Humana Press.Google Scholar
Handelsman, J. (2002). Future trends in biocontrol. In Biological Control of Crop Diseases, ed. Gnanamanickam, S. S., pp. 443448. New York: Dekker.Google Scholar
Hansen, R. W., Spencer, N. R., Fornasari, L., Quimby, P. C., Jr., Pemberton, R. W., & Nowierski, R. M. (2004). Leafy spurge. In Biological Control of Invasive Plants in the United States, ed. Coombs, E. M., Clark, J. K., Piper, G. L. & Cofrancesco, A. F. Jr., pp. 233262. Corvallis: Oregon State University Press.Google Scholar
Harris, J. (2000). Chemical Pesticide Markets, Health Risks, and Residues. Wallingford, UK: CABI Publishing.Google Scholar
Harris, P. (1981). Stress as a strategy in the biological control of weeds. In Biological Control in Plant Production, ed. Papavizas, G. C., pp. 333340. Totowa, NJ: Allanheld.Google Scholar
Harris, P. (1986). Biological control of weeds. In Biological Plant and Health Protection, Biological Control of Plant Pests and of Vectors of Human and Animal Diseases, ed. Franz, J. M., International Symposium of the Akademie der Wissenschaften und der Literatur, Mainz, Nov. 15–17, 1984, pp. 123138. Stuttgart: G. Fischer.Google Scholar
Harris, P. (1991). Classical biocontrol of weeds: Its definitions, selection of effective agents, and administrative-political problems. The Canadian Entomologist, 123, 827849.Google Scholar
Harris, P., Wilkinson, A. T. S., Thompson, L. S., & Neary, M. (1978). Interaction between the cinnabar moth, Tyria jacobaeae L. (Lep., Arctiidae) and ragwort, Senecio jacobaea L. (Compositae) in Canada. Proceedings of the Fourth International Symposium on Biological Control of Weeds, pp. 174180.Google Scholar
Hartley, S. E. & Jones, C. G. (1997). Plant chemistry and herbivory, or why the world is green. In Plant Ecology, ed. Crawley, M. J., pp. 284324, 2nd edn. Oxford: Blackwell.Google Scholar
Harvey, C. H. & Eubanks, M. D. (2005). Intraguild predation of parasitoids by Solenopsis invicta: A non-disruptive interaction. Entomologia Experimentalis et Applicata, 114, 127135.Google Scholar
Havill, N. P., Davis, G., Mausel, D. I., Klein, J., McDonald, R., Jones, C., Fischer, M., Salom, S., & Caccone, A. (2012). Hybridization between a native and introduced predator of Adelgidae: An unintended result of classical biological control. Biological Control, 63, 359369.Google Scholar
Hawkins, B. A. & Marino, P. C. (1997). The colonization of native phytophagous insects in North America by exotic parasitoids. Oecologia, 112, 566571.Google Scholar
Hawkins, B. A., Mills, N. J., Jervis, M. A., & Price, P. W. (1999). Is the biological control of insects a natural phenomenon? Oikos, 86, 493506.Google Scholar
Heimpel, G. E. & Mills, N. J. (2017). Biological Control: Ecology and Applications. Cambridge: Cambridge University Press.Google Scholar
Henry, J. E. & Oma, E. A. (1981). Pest control by Nosema locustae, a pathogen of grasshoppers and crickets. In Microbial Control of Pests and Plant Diseases 1970–1980, ed. Burges, H. D., pp. 573586. New York: Academic Press.Google Scholar
Herren, H. R., Bird, T. J., & Nadel, D. J. (1987). Technology for automated aerial release of natural enemies of the cassava mealybug and cassava green mite. Insect Science and Application, 8, 883885.Google Scholar
Higley, L. G. & Pedigo, L. P. (eds.) (1996). Economic Thresholds for Integrated Pest Management. Lincoln: University of Nebraska Press.Google Scholar
Hintz, W. (2007). Development of Chondrostereum purpureum as a mycoherbicide for deciduous brush control. In Biological Control: A Global Perspective, ed. Vincent, C., Goettel, M. S., & Lazarovits, G., pp. 284292. Wallingford, UK: CABI Publishing.Google Scholar
Hinz, H. L., Schwarzländer, M., & Gaskin, J. (2008). Does phylogeny explain the host-choice behaviour of potential biological control agents for Brassicaceae weeds? In Proceedings, XII International Symposium of Biological Control of Weeds, ed. Julien, M. H., Sforza, R., Bon, M. C., Evans, H. C., Hatcher, P. E., Hinz, H. L., & Rector, B. G., pp. 418425. Wallingford, UK: CABI Publishing.Google Scholar
Hinz, H. L., Schwarzländer, M., Gassmann, A., & Bourchier, R. S. (2014). Successes we may not have had: A retrospective analysis of selected weed biological control agents in the United States. Invasive Plant Science and Management, 7, 565579.Google Scholar
Hoch, H. C. & Fuller, M. S. (1977). Mycoparasitic relationships: I. Morphological features of interactions between Pythium acanthicum and several fungal hosts. Archives of Microbiology, 111, 207224.Google Scholar
Hochberg, M. E. (1989). The potential role of pathogens in biological control. Nature, 337, 262264.Google Scholar
Hoddle, M. S. (1999). Biological control of vertebrate pests. In Handbook of Biological Control: Principles and Applications of Biological Control, ed. Bellows, T. S. & Fisher, T. W., pp. 955975. San Diego, CA: Academic Press.Google Scholar
Hoddle, M. S., Van Driesche, R. G., & Sanderson, J. P. (1998). Biology and use of the whitefly parasitoid Encarsia formosa. Annual Review of Entomology, 43, 645669.Google Scholar
Hoffmann, A. A., Montgomery, B. L., Popovici, J., Iturbe-Ormaetxe, I., Johnson, P. H., Muzzi, F., Greenfield, M., Durkan, M., Leong, Y. S., Dong, Y., et al. (2011). Successful establishment of Wolbachia in Aedes populations to suppress dengue transmission. Nature, 476, 454457.Google Scholar
Hoffmann, J. H. (1995). Biological control of weeds: The way forward, a South African perspective. BCPC Symposium Proceedings, 64, 7789.Google Scholar
Hoffmann, J. H. & Moran, V. C. (1992). Oviposition patterns and the supplementary roles of a seed-feeding weevil, Rhyssomatus marginatus (Coleoptera: Curculionidae), in the biological control of a perennial leguminous weed, Sesbania punicea. Bulletin of Entomological Research, 82, 343347.Google Scholar
Hoffmann, J. H. & Moran, V. C. (1995). Localized failure of a weed biological control agent attributed to insecticide drift. Agriculture, Ecosystems & Environment, 52, 197203.Google Scholar
Hoffmann, M. P. & Frodsham, A. C. (1993). Natural Enemies of Vegetable Insect Pests. Ithaca, NY: Cornell Cooperative Extension.Google Scholar
Hokkanen, H. & Lynch, J. M. (eds.) (1995). Biological Control: Benefits and Risks. Cambridge: Cambridge University Press.Google Scholar
Hokkanen, H. & Pimentel, D. (1984). New approach for selecting biological control agents. Canadian Entomologist, 116, 11091121.Google Scholar
Hokkanen, H. & Pimentel, D. (1986). New associations in biological control: Theory and practice. Canadian Entomologist, 121, 829840.Google Scholar
Holloway, J. K. (1957). Weed control by insect. Scientific American, 197 (1), 5662.Google Scholar
Holm, L. G., Plucknett, D. L., Pancho, J. V., & Herberger, J. P. (1977). The World's Worst Weeds: Distribution and Biology. Honolulu: University of Hawaii Press.Google Scholar
Hough-Goldstein, J., Lake, E., & Reardon, R. (2012). Status of an ongoing biological control program for the invasive vine, Persicaria perfoliata, in eastern North America. BioControl, 57, 181189.Google Scholar
Howarth, F. G. (1983). Biological control: Panacea or Pandora's box? Proceedings of the Hawaiian Entomological Society, 24, 239244.Google Scholar
Howarth, F. G. (1991). Environmental impacts of classical biological control. Annual Review of Entomology, 36, 485510.Google Scholar
Hoy, M. A. (1993). Biological control in U.S. agriculture: Back to the future. American Entomologist, 39, 140150.CrossRefGoogle Scholar
Hrusa, G. F. & Gaskin, J. F. (2008). The Salsola complex in California (Chenopodiaceae): Characterization and status of Salsola australis and the autochthonous allopolyploid Salsola ryanii sp. nov. Madroño, 55, 113131.Google Scholar
Huang, N., Enkegaard, A., Osborne, L. S., Ramakers, P. M. J., Messelink, G. J., Pijnakker, J., & Murphy, G. (2011) The Banker Plant Method in biological control. Critical Reviews in Plant Sciences, 30, 3, 259278Google Scholar
Huffaker, C. B. (1958). Experimental studies on predation: Dispersion factors and predator–prey oscillations. Hilgardia, 27, 343383.Google Scholar
Huffaker, C. B., Messenger, P. S., & DeBach, P. (1971). The natural enemy component in natural control and the theory of biological control. In Biological Control, ed. Huffaker, C. B., pp. 1667. New York: Plenum Press.Google Scholar
Huffaker, C. B., Shea, K. P., & Herman, S. G. (1963). Experimental studies on predation: Complex dispersion and levels of food in an acarine predator-prey interaction. Hilgardia, 34, 305330.Google Scholar
Hughes, C. E. (1995). Protocols for plant introductions with particular reference to forestry: Changing perspectives on risks to biodiversity and economic development. Proceedings Brighton Crop Protection Conference Symposium: Weeds in a Changing World, 64, 1532.Google Scholar
Huigens, M. E., Woelke, J. B., Pashalidou, F. G., Bukovinszky, T., Smid, H. M., & Fatouros, N. E. (2010). Chemical espionage on species-specific butterfly anti-aphrodisiacs by hitchhiking Trichogramma wasps. Behavioral Ecology, 21, 470478.Google Scholar
Hull, L. A., Hickey, K. D., & Kanour, W. W. (1983). Pesticide usage patterns and associated pest damage in commercial apple orchards of Pennsylvania. Journal of Economic Entomology, 76, 577583.Google Scholar
Hunter-Fujita, F. R., Entwistle, P. F., Evans, H. F., & Crook, N. E. (1998). Insect Viruses and Pest Management. Chichester, UK: John Wiley & Sons.Google Scholar
Jackson, T. A., Pearson, J. F., O'Callaghan, M., Mahanty, H. K., & Willocks, M. J. (1992). Pathogen to product-development of Serratia entomophila (Enterobacteriaceae) as a commercial biological control agent for the New Zealand grass grub (Costelytra zealandica). In Use of Pathogens in Scarab Pest Management, ed. Glare, T. R. & Jackson, T. A., pp. 191198. Andover, UK: Intercept Ltd.Google Scholar
James, D. G., Orre-Gordon, S., Reynolds, O. L., & Simpson, M. (2012). Employing chemical ecology to understand and exploit biodiversity for pest management. In Biodiversity and Insect Pests: Key Issues for Sustainable Management, ed. Gurr, G. M., Wratten, S. D., Snyder, W. E., & Read, D. M. Y., pp. 185195. Chichester, UK: Wiley-Blackwell.Google Scholar
Jeschke, J. M., Gómez Aparicio, L., Haider, S., Heger, T., Lortie, C. J., Pyšek, P., & Strayer, D. (2012). Support for major hypotheses in invasion biology is uneven and declining. Neobiota, 14, 120.Google Scholar
Johnson, M. T., Follett, P. A., Taylor, A. D., & Jones, V. P. (2005). Impacts of biological control and invasive species on a non-target native Hawaiian insect. Oecologia, 142, 529540.Google Scholar
Julien, M., Center, T. D., & Tipping, P. W. (2002). Floating fern (Salvinia). In Biological Control of Invasive Plants in the Eastern United States, ed. Van Driesche, R., Blossey, B., Hoddle, M., Lyon, S., & Reardon, R., pp. 1732. USDA Forest Service, FHTET-2002-04.Google Scholar
Julien, M. & White, G. (1997). Biological Control of Weeds: Theory and Practical Application (ACIAR Monograph 49). Canberra: Australian Centre for International Agricultural Research.Google Scholar
Kairo, M. T. K., Paraiso, O., Das Gautam, R., & Peterkin, D. D. (2013). Cryptolaemus montrouzieri (Mulsant) (Coccinellidae: Scymninae): A review of biology, ecology, and use in biological control with particular references to potential impact on non-target organisms. CAB Reviews, 8, 120.Google Scholar
Kaplan, I. & Thaler, J. S. (2010). Plant resistance attenuates the consumptive and non-consumptive impacts of predators on prey. Oikos, 109, 11051113.Google Scholar
Kenis, M., Auger-Rozenberg, M. A., Roques, A., Timms, L., Péré, C., Cock, M. J. W., Settele, J., Augustin, S., & Lopez-Vaamonde, C. (2009). Ecological effects of invasive alien insects. Biological Invasions, 11, 2145.Google Scholar
Kenis, M., Hurley, B., Hajek, A. E., & Cock, M. (2017). Classical biological control against insect tree pests: Facts and figures. Biological Invasions, 19, 34013417.Google Scholar
Kerr, A. (1980). Biological control of crown gall through production of agrocin 84. Plant Disease, 64, 2530.Google Scholar
Kerry, B. R. (2000). Rhizosphere interactions and the exploitation of microbial agents for the biological control of plant-parasitic nematodes. Annual Review of Phytopathology, 38, 423441.Google Scholar
Kerry, B. R. (2001). Exploitation of the nematophagous fungal Verticillium chlamydosporium Goddard for the biological control of root-knot nematodes (Meloidogyne spp.). In Fungi as Biocontrol Agents: Progress, Problems and Potential, ed. Butt, T. M., Jackson, C., & Magan, N., pp. 155167. Wallingford, UK: CAB International.Google Scholar
Kessler, A. & Baldwin, I. T. (2001). Defensive function of herbivore-induced plant volatile emission in nature. Science, 291, 21412144.Google Scholar
Kimberling, D. N. (2004). Lessons from history: Predicting successes and risks of intentional introductions for arthropod biological control. Biological Invasions 6, 301318.Google Scholar
Kiritani, K., Kawahara, S., Sasaba, T., & Nakasuji, F. (1972). Quantitative evaluation of predation by spiders on the green rice leafhopper, Nephotettix cincticeps Uhler, a sight count method. Researches on Population Ecology, 13, 187200.Google Scholar
Klemola, N., Andersson, T., Ruohomäki, K., & Klemola, T. (2010). Experimental test of parasitism hypothesis for population cycles of a forest lepidopteran. Ecology, 91, 25062513.Google Scholar
Kogan, M. (1994). Plant resistance in pest management. In Introduction to Insect Pest Management, ed. Metcalf, R. L. & Luckmann, W. H., 3rd edn., pp. 73128. New York: John Wiley & Sons.Google Scholar
Kogan, M. (1998). Integrated pest management: Historical perspectives and contemporary developments. Annual Review of Entomology, 43, 243270.Google Scholar
Kok, L. T. (2001). Classical biological control of nodding and plumeless thistles. Biological Control, 21, 206213.Google Scholar
Kondo, A. & Hiramatsu, T. (1999). Resurgence of the peach silver mite, Aculus fockeui (Nalepa et Trouessart) (Acari: Eriophyidae), induced by a synthetic pyrethroid fluvalinate. Applied Entomology and Zoology, 34, 531535.Google Scholar
Krebs, C. J. (2001). Ecology: The Experimental Analysis of Distribution and Abundance, 5th edn. San Francisco, CA: Benjamin Cummings.Google Scholar
Krieg, A., Huger, A. M., Langenbruch, G. A., & Schnetter, W. (1983). Bacillus thuringiensis var. tenebrionis: A new pathotype effective against larvae of Coleoptera. Zeitschrift für Angewandte Entomologie, 96, 500508.Google Scholar
Kuris, A. M. (2003). Did biological control cause extinction of the coconut moth, Levuana iridescens, in Fiji? Biological Invasions, 5, 133141.Google Scholar
Lafferty, K. D. & Kuris, A. M. (1996). Biological control of marine pests. Ecology, 77, 19892000.Google Scholar
Landis, D. A. (2017). Designing agricultural landscapes for biodiversity-based ecosystem services. Basic and Applied Ecology, 18, 112.Google Scholar
Landis, D. A., Wratten, S. D., & Gurr, G. M. (2000). Habitat management to conserve natural enemies of arthropod pests in agriculture. Annual Review of Entomology, 45, 175201.Google Scholar
Langewald, J. & Kooyman, C. (2007). Green Muscle, a fungal biopesticide for control of grasshoppers and locusts in Africa. In Biological Control: A Global Perspective, ed. Vincent, C., Goettel, M. S., Lazarovits, G., pp. 311318. Wallingford, UK: CABI Publishing.Google Scholar
Larson, D. L., James, J. B., Grace, B., & Larson, J. L. (2008). Long-term dynamics of leafy spurge (Euphorbia esula) and its biocontrol agent, flea beetles in the genus Aphthona. Biological Control, 47, 250256.Google Scholar
Lawton, J. H. (1990). Biological control of plants: A review of generalisations, rules and principles using insects as agents. In Alternatives to the Chemical Control of Weeds: Proceedings of International Conference held at Rotorua, New Zealand, July 1989, ed. Bassett, C., Whitehouse, L. J., & Zabkiewicz, J. A., 317. Rotorua, New Zealand: Ministry of Forestry, Forest Research Institute.Google Scholar
Lawton, J. H. & McNeill, S. (1979). Between the devil and the deep blue sea: On the problem of being an herbivore. In Population Dynamics, ed. Anderson, K., Turner, B., & Taylor, L. R., pp. 223245. Oxford: Blackwell.Google Scholar
Leake, D. V, Leake, J. B., & Roeder, M. L. (1993). Desert and Mountain Plants of the Southwest. Norman: Oklahoma University Press.Google Scholar
Lennox, C. L., Morris, M. J., Van Rooi, C., Serdani, M., Wood, A. R., Den Breeÿen, A., Markram, J. L., & Samuels, G. (2004). A decade of biological control of Acacia saligna in South Africa, using the gall rust fungus, Uromycladium tepperianum. Proceedings of the XI International Symposium on Biological Control of Weeds, Canberra, Australia, April 27–May 2, 2003, pp. 574575.Google Scholar
Letourneau, D. K., Jedlicka, J. A., Bothwell, S. B., & Moreno, C. R. (2009). Effects of natural enemy biodiversity on the suppression of arthropod herbivores in terrestrial ecosystems. Annual Review of Ecology, Evolution and Systematics, 40, 573592.Google Scholar
Lewis, S. (1989). Cane Toads: An Unnatural History. New York: Dolphin/Doubleday.Google Scholar
Lewis, W. J., van Lenteren, J. C., Phatak, S. C., & Tumlinson, J. H., III (1997). A total system approach to sustainable pest management. Proceedings of the National Academy of Sciences, U.S.A., 94, 12,24312,248.Google Scholar
Li, L.-Y. (1994). Worldwide use of Trichogramma for biological control of different crops: A survey. In Biological Control with Egg Parasitoids, ed. Wajnberg, E. & Hassan, S. A., pp. 3753. Wallingford, UK: CABI Publishing.Google Scholar
Li, Z., Alves, S. B., Roberts, D. W., Fan, M., Delalibera, I., Jr., Tang, J., et al. (2010). Biological control of insects in Brazil and China: History, current programs and reasons for their successes using entomopathogenic fungi. Biocontrol Science and Technology, 20, 117136.Google Scholar
Liang, W. & Huang, M. (1994). Influence of citrus orchard ground cover plants on arthropod communities in China: A review. Agriculture, Ecosystems & Environment, 45, 175201.Google Scholar
Lindow, S. E. & Brandl, M. T. (2003). Microbiology of the phyllosphere. Applied and Environmental Microbiology, 69, 18751883.Google Scholar
Lomer, C. J., Bateman, R. P., Johnson, D. L., Langewald, J., & Thomas, M. (2001). Biological control of locusts and grasshoppers. Annual Review of Entomology, 46, 667702.Google Scholar
Lomer, C. J., Thomas, M. B., Douro-Kpindou, O.-K., Gbongboui, C., Godonou, I., Langewald, J., & Shah, P. A. (1997). Control of grasshoppers, particularly Hieroglyphus daganensis, in northern Benin using Metarhizium anisopliae. Memoirs of the Entomological Society of Canada, 171, 301311.Google Scholar
Losey, J. E. & Denno, R. F. (1998). Positive predator–predator interactions: Enhanced predation rates and synergistic suppression of aphid populations. Ecology, 79, 21432152.Google Scholar
Louda, S. M., Kendall, D., Connor, J., & Simberloff, D. (1997). Ecological effects of an insect introduced for the biological control of weeds. Science, 277, 10881090.Google Scholar
Louda, S. M., Pemberton, R. W., Johnson, M. T., & Follett, P. A. (2003). Non-target effects – The Achilles’ heel of biological control? Annual Review of Entomology, 48, 365396.Google Scholar
Luck, R. F., Shepard, B. M., & Kenmore, P. E. (1988). Experimental methods for evaluating arthropod natural enemies. Annual Review of Entomology, 33, 367391.Google Scholar
Luckmann, W. H. & Metcalf, R. L. (1994). The pest management concept. In Introduction to Insect Pest Management, ed. Metcalf, R. L. & Luckmann, W. H., 3rd edn., pp. 134. New York: John Wiley & Sons.Google Scholar
Lumsden, R. D., Lewis, J. A., & Locke, J. C. (1993). Managing soilborne plant pathogens with fungal antagonists. In Pest Management: Biologically Based Technologies, ed. Lumsden, R. D. & Vaughn, J. L., pp. 196203. Washington, DC: American Chemical Society.Google Scholar
Lym, R. G. & Zollinger, R. K. (1995). Integrated management of leafy spurge. North Dakota State Extension Service Publication W-866.Google Scholar
Lynch, L. D. & Thomas, M. B. (2000). Nontarget effects in the biocontrol of insects with insects, nematodes and microbial agents: The evidence. Biocontrol News & Information, 21, 117N130N.Google Scholar
Maczey, N., Tanner, R., & Shaw, R. (2012). Understanding and addressing the impact of invasive non-native species in the UK Overseas Territories in the South Atlantic: A review of the potential for biocontrol/ Preliminary results Ascension Island. Report ref # TR10086. Wallingford, UK: CABI Publishing.Google Scholar
Madden, J. L. (1968). Behavioural responses of parasites of the symbiotic fungus associated with Sirex noctilio F. Nature, 218, 189190.Google Scholar
Malcolm, S. B. (1992). Prey defence and predator foraging. In Natural Enemies: The Population Biology of Predators, Parasites and Diseases, ed. Crawley, M. J., pp. 458475. Oxford: Blackwell Scientific Publications.Google Scholar
Marshall, J. D. & Fenner, F. (1958). Studies in the epidemiology of infectious myxomatosis of rabbits: VI. Changes in the innate resistance of Australian wild rabbits exposed to myxomatosis. Journal of Hygiene, 56, 288302.Google Scholar
Mathre, D. E., Cook, R. J., & Callan, N. W. (1999). From discovery to use: Traversing the world of commercializing biocontrol agents for plant disease control. Plant Disease, 83, 972983.Google Scholar
McCoy, C. W., Stuart, R. J., Duncan, L. W., & Shapiro-Ilan, D. I. (2007). Application and evaluation of entomopathogens for citrus pest control. In Field Manual of Techniques in Invertebrate Pathology, ed. Lacey, L. A. & Kaya, H. K., pp. 567581. Dordrecht, Netherlands: Springer.Google Scholar
McEvoy, P. B. & Coombs, E. M. (1999). Biological control of plant invaders: Regional patterns, field experiments, and structured population models. Ecological Applications, 9, 387401.Google Scholar
McEvoy, P. B., Rudd, N. T., Cox, C. S., & Huso, M. (1993). Disturbance, competition, and herbivory effects on ragwort Senecio jacobaea populations. Ecological Monographs, 63, 5575.Google Scholar
McFadyen, R. E. C. (1998). Biological control of weeds. Annual Review of Entomology, 43, 369393.Google Scholar
McFadyen, R. E. C. (2000). Successes in biological control of weeds. In Proceedings of the X International Symposium on Biological Control of Weeds, ed. Spencer, N. R., pp. 314. Bozeman: Montana State University.Google Scholar
Meehan, T. D., Werling, B. P., Landis, D. A., & Gratton, C. (2011). Agricultural landscape simplification and insecticide use in the Midwestern United States. Proceedings of the National Academy of Sciences, U.S.A., 108, 11,50011,505.Google Scholar
Mendes, R., Kruijt, M., De Bruijn, I., Dekkers, E., Van der Voort, M., Schneider, J. H. M., Piceno, Y. M., DeSantis, T. Z., Andersen, G. L., Bakker, P. A., et al. (2011). Deciphering the rhizosphere microbiome for disease-suppressive bacteria. Science, 332, 10971100.Google Scholar
Messelink, G. J., Bennison, J., Alomar, O., Ingegno, B. L., Tavella, L., Shipp, L., Palevsky, E., & Wäckers, F. L. (2014). Approaches to conserving natural enemy populations in greenhouse crops: Current methods and future prospects. BioControl, 59, 377393.Google Scholar
Milbrath, L. R. & Nechols, J. R. (2014). Plant-mediated interactions: Considerations for agent selection in weed biological control programs. Biological Control, 72, 8090.Google Scholar
Mills, N. J. (2006). Accounting for differential success in the biological control of homopteran and lepidopteran pests. New Zealand Journal of Ecology, 30, 6172.Google Scholar
Moosavi, M. R. & Zare, R. (2015). Factors affecting commercial success of biocontrol agents of phytonematodes. In Biocontrol Agents of Phytonematodes, ed. Askary, T. & Martinelli, P. R. P., pp. 423445. Wallingford, UK: CABI.Google Scholar
Mora, C., Tittensor, D. P., Adl, S., Simpson, A. G. B., & Worm, B. (2011). How many species are there on earth and in the ocean? PLoS Biology, 9, e1001127.Google Scholar
Morrison, L. W., Dall'aglio-Holvercem, C. G., & Gilbert, L. E. (1997). Oviposition behavior and development of Pseudacteon flies (Diptera, Phoridae), parasitoids of Solenopsis fire ants (Hymenoptera, Formicidae). Environmental Entomology, 26, 716724.Google Scholar
Moscardi, F. (1999). Assessment of the application of baculoviruses for control of Lepidoptera. Annual Review of Entomology, 44, 257289.Google Scholar
Mota-Sanchez, D., Bills, P. S., & Whalon, M. E. (2002). Arthropod resistance to pesticides: Status and overview. In Pesticides in Agriculture and the Environment, ed. Wheeler, W. B., pp. 241272. New York: Dekker.Google Scholar
Mota-Sanchez, D. & Wise, J. C. (2017). Arthropod Pesticide Resistance Database 2016. Retrieved from www.pesticideresistance.orgGoogle Scholar
Murdoch, W. & Briggs, C. J. (1996). Theory for biological control: Recent developments. Ecology, 77, 20012013.Google Scholar
Murdoch, W., Briggs, C. J., & Swarbrick, S. (2005). Host suppression and stability in a parasitoid–host system: Experimental demonstration. Science, 309, 610613.Google Scholar
Murphy, B. C., Rosenheim, J. A., Granett, J., Pickett, C. H., & Dowell, R. V. (1998). Measuring the impact of a natural enemy refuge: The prune tree/vineyard example. In Enhancing Biological Control: Habitat Management to Promote Natural Enemies of Agricultural Pests, ed. Pickett, C. H. & Bugg, R. L., pp. 297309. Berkeley: University of California Press.Google Scholar
Murphy, G. D., Ferguson, G., Frey, K., Lambert, L., Mann, M., & Matteoni, J. (2002). The use of biological control in Canadian greenhouse crops. In Integrated Control in Protected Crops, Temperate Climate, ed. Enkegaard, S.. IOBC WPRS Bulletin, 25(1), 193196.Google Scholar
Murray, E. (1993). The sinister snail. Endeavour, 17, 7883.Google Scholar
Myers, J. H. (1985). How many insect species are necessary for successful biocontrol of weeds? In Proceedings of the 6th International Symposium on the Biological Control of Weeds, Agriculture Canada, ed. Delfosse, E. S., pp. 7782. Ottawa, Canada: Canadian Govt. Printing Office.Google Scholar
Myers, J. H. & Harris, P. (1980). Distribution of Urophora galls in flowers heads of diffuse and spotted knapweed in British Columbia. Journal of Applied Ecology, 17, 359367.Google Scholar
Myers, J. H. & Risley, C. (2000). Why reduced seed production is not necessarily translated into successful biological weed control. In Proceedings, X International Symposium on Biological Control of Weeds, ed. Spencer, N. R., pp. 569581. Bozeman: Montana State University.Google Scholar
Naranjo, S. E. & Ellsworth, P. C. (2009). Fifty years of the integrated control concept: Moving the model and implementation forward in Arizona. Pest Management Science, 65, 12671286.Google Scholar
Naranjo, S. E. & Ellsworth, P. C. (2010). Fourteen years of Bt cotton advances IPM in Arizona. Southwestern Entomologist, 35, 437444.Google Scholar
National Academy of Sciences (US). (1988). Research Briefings 1987: Report of the Research Panel on Biological Control in Managed Ecosystems. Washington, DC: National Academy Press.Google Scholar
National Research Council (US), Committee on Pest and Pathogen Control Through Management of Biological Control Agents and Enhanced Cycles and Natural Processes. (1996). Ecologically Based Pest Management: New Solutions for a New Century. Washington, DC: National Academy Press.Google Scholar
Neuenschwander, P. & Herren, H. (1988). Biological control of the cassava mealybug, Phenacoccus manihoti, by the exotic parasitoid Epidinocarsis lopezi in Africa. Philosophical Transactions of the Royal Society of London B, 318, 319333.Google Scholar
Newhouse, J. R. (1990). Chestnut blight. Scientific American, 263(1), 106111.Google Scholar
Newman, R. M., Thompson, D. C., & Richman, D. B. (1998). Conservation strategies for the biological control of weeds. In Conservation Biological Control, ed. Barbosa, P., pp. 371396. San Diego, CA: Academic Press.Google Scholar
Nielsen, C., Milgroom, M. G., & Hajek, A. E. (2005). Genetic diversity in the gypsy moth fungal pathogen Entomophaga maimaiga from founder populations in North America and source populations in Asia. Mycological Research, 109, 941950.Google Scholar
Nilsson, C. (1985). Impact of ploughing on emergence of pollen beetle parasitoids after hibernation. Zeitschrift für Angewandte Entomologie, 100, 302308.Google Scholar
Nordlund, D. A. (1996). Biological control, integrated pest management and conceptual models. Biocontrol News & Information, 17, 35N44N.Google Scholar
Nottingham, L. B., Dively, G. P., Schultz, P. B., Herbert, D. A., & Kuhar, T. P. (2016). Natural history, ecology, and management of the Mexican bean beetle (Coleoptera: Coccinellidae) in the United States. Journal of Integrated Pest Management, 7(1), 2.Google Scholar
Nowierski, R. M., Zeng, Z., Schroeder, D., Gassmann, A., FitzGerald, B. C., & Cristofaro, M. (2002). Habitat associations of Euphorbia and Aphthona species for Europe: Development of predictive models for natural enemy release with ordination analysis. Biological Control, 23, 117.Google Scholar
Nyrop, J., English-Loeb, G., & Roda, A. (1998). Conservation biological control of spider mites in perennial cropping systems. In Conservation Biological Control, ed. Barbosa, P., pp. 307333. San Diego, CA: Academic Press.Google Scholar
Oliver, K. M., Noge, K., Huang, E. M., Campos, J. M., Becerra, J. X., & Hunter, M. S. (2012). Parasitic wasp responses to symbiont-based defense in aphids. BMC Biology, 10, 11.Google Scholar
Onstad, D. W., Fuxa, J. R., Humber, R. A., Oestergaard, J., Shapiro-Ilan, D. I., Gouli, V. V., Anderson, R. S., Andreadis, T. G., & Lacey, L. A. (2006). An Abridged Glossary of Terms Used in Invertebrate Pathology, 3rd edn. Society for Invertebrate Pathology. Retrieved from www.sipweb.org/resources/glossary.html.Google Scholar
Orr, D. (2009). Biological control and integrated pest management. In Integrated Pest Management: Innovation-Development Processes, ed. Peshin, R. & Dhawan, A. K., pp. 207239. Dordrecht, Netherlands: Springer.Google Scholar
OTA (Office of Technology Assessment; US Congress). (1995). Biologically Based Technologies for Pest Control, OTA-ENV-636. Washington, DC: US Government Printing Office.Google Scholar
Paine, R. T. (1996). Preface. In Food Webs: Integration of Patterns and Dynamics, ed. Polis, G. A. & Winemille, K. O., pp. ixx. New York: Chapman & Hall.Google Scholar
Parra, J. R. P. (2014). Biological control in Brazil: An overview. Scientia Agricola, 71, 345355.Google Scholar
Parrella, M. P., Hansen, L. S., & van Lenteren, J. (1999). Glasshouse environments. In Handbook of Biological Control, ed. Bellows, T. S. & Fisher, T. W., pp. 819839. San Diego, CA: Academic Press.Google Scholar
Pedigo, L. P. (1996). Entomology and Pest Management, 2nd edn. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Pemberton, R. W. (2000). Predictable risk to native plants in weed biological control. Oecologia, 125, 489494.Google Scholar
Perfecto, I. & Castiñeiras, A. (1998). Deployment of the predaceous ants and their conservation in agroecosystems. In Conservation Biological Control, ed. Barbosa, P., pp. 269289. San Diego, CA: Academic Press.Google Scholar
Perkins, J. H. & Garcia, R. (1999). Social and economic factors affecting research and implementation of biological control. In Handbook of Biological Control, ed. Bellows, T. S. & Fisher, T. W., pp. 9931009. San Diego, CA: Academic Press.Google Scholar
Perkins, R. C. L. (1897). The introduction of beneficial insects into the Hawaiian Islands. Nature, 55, 499500.Google Scholar
Pickett, C. H. & Bugg, R. L. (eds.) (1998). Enhancing Biological Control: Habitat Management to Promote Natural Enemies of Agricultural Pests. Berkeley: University of California Press.Google Scholar
Pimentel, D. (ed.) (2002). Biological Invasions: Economic and Environmental Costs of Alien Plant, Animal, and Microbe Species. Boca Raton, FL: CRC Press.Google Scholar
Pimentel, D., Acquay, H., Biltonen, M., Rice, P., Silva, M., Nelson, J., Lipner, V., Giordano, S., Horowitz, A., & D'Amore, M. (1992). Assessment of environmental and economic impacts of pesticide use. In The Pesticide Question: Environment, Economics, and Ethics, ed. Pimentel, D. & Lehman, H., pp. 4784. New York: Chapman & Hall.Google Scholar
Price, P. W. (1984). The concept of the ecosystem. In Ecological Entomology, ed. Huffaker, C. B. & Rabb, R. L., pp. 1950. New York: John Wiley & Sons.Google Scholar
Quarles, W. (ed.) (2015). 2015 Directory of least-toxic pest control products. The IPM Practitioner, 34 (11/12), 141.Google Scholar
Quicke, D. L. J. (1997). Parasitic Wasps. London: Chapman & Hall.Google Scholar
Radcliffe, E. B. & Flanders, K. L. (1998). Biological control of alfalfa weevil in North America. Integrated Pest Management Reviews, 3, 225242.Google Scholar
Ramula, S., Knight, T. M., Burns, J. H., & Buckley, Y. M. (2008). General guidelines for invasive plant management based on comparative demography of invasive and native plant populations. Journal of Applied Ecology 45, 11241133.Google Scholar
Reilly, J. R. & Hajek, A. E. (2012). Prey processing by avian predators increases virus transmission in the gypsy moth. Oikos, 121, 13111316.Google Scholar
Rezende, J. M., Beatriz, A., Zanardo, R., da Silva Lopes, M., Delalibera, I., Jr., & Rehner, S. A. (2015). Phylogenetic diversity of Brazilian Metarhizium associated with sugarcane agriculture. BioControl, 60, 495505.Google Scholar
Rishbeth, J. R. (1975). Stump inoculation: A biological control of Fomes annosus. In Biology and Control of Soil-borne Pathogens, ed. Bruehl, G. W., pp. 158162. St. Paul, MN: American Phytopathological Society.Google Scholar
Root, R. B. (1973). Organization of plant-arthropod association in simple and diverse habitats: The fauna of collards (Brassica oleracea). Ecological Monographs, 43, 95124.Google Scholar
Rosenheim, J. A. (1998). Higher-order predators and the regulation of insect herbivore populations. Annual Review of Entomology, 43, 421447.Google Scholar
Rosskopf, E. N., Charudattan, R., & Kadir, J. B. (1999). Use of plant pathogens in weed control. In Handbook of Biological Control, ed. Bellows, T. S. & Fisher, T. W., pp. 891918. San Diego, CA: Academic Press.Google Scholar
Ruberson, J. R., Kring, T. J., & Elkassabany, N. (1998). Overwintering and the diapause syndrome of predatory Heteroptera. In Predatory Heteroptera in Agroecosystems: Their Biology and Use in Biological Control, ed. Coll, M. & Ruberson, J. R., pp. 4669. Lanham, MD: Entomological Society of America.Google Scholar
Rutz, D. A. & Watson, D. W. (1998). Parasitoids as a component in an integrated fly-management program on dairy farms. In Mass-Reared Natural Enemies: Application, Regulation, and Needs, ed. Ridgway, R. L., Hoffmann, M. P., Inscoe, M. N., & Glenister, C. S., pp. 185201. Lanham, MD: Entomological Society of America.Google Scholar
Samuels, G. J. & Hebbar, P. K. (2015). Trichoderma: Identification and Agricultural Applications. St. Paul, MN: APS Press.Google Scholar
Sasser, J. N. & Freckman, D. W. (1987). A world perspective on nematology: The role of the society. In Vistas on Nematology: A Commemoration of the Twenty-fifth Anniversary of the Society of Nematologists, ed. Veech, J. A. & Dickson, D. W., pp. 714. Hyattsville, MD: Society of Nematologists.Google Scholar
Saunders, G., Cooke, B., McColl, K., Shine, R., & Peacock, T. (2010). Modern approaches for the biological control of vertebrate pests: An Australian perspective. Biological Control, 52, 288295.Google Scholar
Schaab, R. (1996). Economy and ecology of biological control activities in Africa: A case study on the cassava mealybug, Phenacoccus manihoti Mat. Ferr. Dissertation, University Hohenheim, Germany.Google Scholar
Schellhorn, N. A., Gagic, V., & Bommarco, R. (2015). Time will tell: Resource continuity bolsters ecosystem services. Trends in Ecology & Evolution, 30, 524530.Google Scholar
Schlinger, E. I. & Dietrick, E. J. (1960). Biological control of insect pests aided by strip-farming alfalfa in experimental program. California Agriculture, 14(1), 89, 15.Google Scholar
Schmidt, J. M. (1992). Host recognition and acceptance by Trichogramma. In Biological Control with Egg Parasitoids, ed. Wajnberg, E. & Hassan, S. A., pp. 165200. Wallingford, UK: CAB International.Google Scholar
Schroeder, D. (1983). Biological control of weeds. In Recent Advances in Weed Control, ed. Fletcher, W. E., pp. 4178. Farnham Royal, UK: Commonwealth Agricultural Bureau.Google Scholar
Seebens, H., Blackburn, T. M., Dyer, E. E., Genovesi, P., Hulme, P. E., Jeschke, J. M., Pagad, S., Pyšek, P., Winter, M., Arianoutsou, M., et al. (2017). No saturation in the accumulation of alien species worldwide. Nature Communications, 8, 14435.Google Scholar
Shapiro-Ilan, D. I., Hazir, S., & Glazer, I. (2017). Basic and applied research: Entomopathogenic nematodes. In Microbial Agents for Control of Insect Pests: From Discovery to Commercial Development and Use, ed. Lacey, L. A., pp. 91105. Amsterdam: Academic Press.Google Scholar
Shaw, S. R. (1993). Observations on the ovipositional behaviour of Neoneurus mantis, an ant-associated parasitoid from Wyoming (Hymenoptera: Braconidae). Journal of Insect Behavior, 6, 649658.Google Scholar
Shu, S., Swedenborg, P. D., & Jones, R. L. (1990). A kairomone for Trichogramma nubilale (Hymenoptera: Trichogrammatidae): Isolation, identification, and synthesis. Journal of Chemical Ecology, 16, 521529.Google Scholar
Shuler, M. L., Hammer, D. A., Granados, R. R., & Wood, H. A. (1995). Overview of baculovirus-insect culture system. In Baculovirus Expression Systems and Biopesticides, ed. Shuler, M. L., Wood, H. A., Granados, R. R., & Hammer, D. A., pp. 111. New York: John Wiley & Sons.Google Scholar
Silvestri, F. (1906). Contribuzioni alla conoscenza biologica degli imenotteri parassiti: I. Biologica del Litomastix truncatellus (Dalm.). Bollettino del Laboratorio di Zoologia Generale e Agraria della Facolta Agraria in Portici, 1, 1764.Google Scholar
Sivasubramaniam, V. & Subramaniam, S. (2015). Area-wide releases and evaluation of the parasitoid Eretmocerus hayati (Hymenoptera:Aphelinidae) for silverleaf whitefly control. Acta Horticulturae, 1105, 8188.Google Scholar
Slippers, B., de Groot, P., & Wingfield, M. J. (eds.) (2012). The Sirex Woodwasp and its Fungal Symbiont: Research and Management of a Worldwide Invasive Pest. Dordrecht, Netherlands: Springer.Google Scholar
Slobodkin, L. B. (1988). Intellectual problems of applied ecology. BioScience, 38, 337342.Google Scholar
Smart, G. C., Jr. (1995). Entomopathogenic nematodes for the biological control of insects. Journal of Nematology, 27, 529534.Google Scholar
Smith, H. S. (1919). On some phases of insect control by the biological method. Journal of Economic Entomology, 12, 288292.Google Scholar
Sprenkel, R. K., Brooks, W. M., Van Duyn, J. W., & Deitz, L. L. (1979). The effects of three cultural variables on the incidence of Nomuraea rileyi, a phytophagous Lepidoptera, and their predators on soybeans. Environmental Entomology, 3, 334339.Google Scholar
Steinhaus, E. A. (1975). Disease in a Minor Chord. Columbus: Ohio University Press.Google Scholar
Stern, V. M., Smith, R. F., van den Bosch, R., & Hagen, K. S. (1959). The integration of chemical and biological control of the spotted alfalfa aphid: I. The integrated control concept. Hilgardia, 29, 81101.Google Scholar
Stiling, P. (1990). Calculating the establishment rates of parasitoids in classical biological control. American Entomologist, 36, 225230.Google Scholar
Stiling, P. (1993). Why do natural enemies fail in classical biological control programs? American Entomologist, 39, 3137.Google Scholar
Stiling, P. & Cornelissen, T. (2005). What makes a successful biocontrol agent? A meta-analysis of biological control agent performance. Biological Control, 34, 236246.Google Scholar
Stokstad, E. (2013). Pesticide planet. Science, 341, 730731.Google Scholar
Strand, M. R. & Pech, L. L. (1995). Immunological basis for compatibility in parasitoid–host relationships. Annual Review of Entomology, 40, 3156.Google Scholar
Strong, D. R., Lawton, J. H., & Southwood, T. R. E. (1984). Insects on Plants. Oxford: Oxford University Press.Google Scholar
Suckling, D. M. & Sforza, R. F. H. (2014). What magnitude are observed non-target impacts from weed biocontrol? PLoS ONE, 9(1), e84847.Google Scholar
Sutton, J. C. & Peng, G. (1993). Manipulation and vectoring of biocontrol organisms to manage foliage and fruit diseases in cropping systems. Annual Review of Phytopathology, 31, 473493.Google Scholar
Swezey, O. H. (1943). Biographical sketch of the work of Albert Koebele in Hawaii. Bulletin of the Hawaiian Sugar Planters’ Association Experiment Station, Entomological Series Bulletin, 22, 58.Google Scholar
Symondson, W. O. C., Sunderland, K. D., & Greenstone, M. H. (2002). Can generalist predators be effective biocontrol agents? Annual Review of Entomology, 47, 561594.Google Scholar
Takken, W. & Knols, B. G. J. (2009). Malaria vector control: Current and future strategies. Trends in Parasitology, 25, 101104.Google Scholar
Tanada, Y. & Kaya, H. K. (1993). Insect Pathology. San Diego, CA: Academic Press.Google Scholar
Tang, W. H. (1994). Yield-increasing bacteria (YIB) and biocontrol of sheath blight of rice. In Improving Plant Productivity and Rhizosphere Bacteria, ed. Ryder, M. H., Stephens, P. M., & Bowen, G. D., pp. 267273. Adelaide, Australia: CSIRO, Division of Soils.Google Scholar
Tauber, M. J., Tauber, C. A., Daane, K. M., & Hagen, K. S. (2000). Commercialization of predators: Recent lessons from green lacewings (Neuroptera: Chrysopidae: Chrysoperla). American Entomologist, 46, 2638.Google Scholar
Thacker, J. R. M. (2002). An Introduction to Arthropod Pest Control. Cambridge: Cambridge University Press.Google Scholar
Thies, C. & Tscharntke, T. (1999). Landscape structure and biological control in agroecosystems. Science, 285, 893895.Google Scholar
Thomas, M. B. (1999). Ecological approaches and the development of “truly integrated” pest management. Proceedings of the National Academy of Sciences, U.S.A., 96, 59445951.Google Scholar
Thomas, M. B. (2018). Biological control of human disease vectors: A perspective on challenges and opportunities. BioControl, 63(1), 6169.Google Scholar
Thomas, M. B., Godfray, H. C. J., Read, A. F., van den Berg, H., Tabashnik, B. E., van Lenteren, J. C., Waage, J. K., & Takken, W. (2012). Lessons from agriculture for the sustainable management of malaria vectors. PLOS Medicine, 9(7), e1001262.Google Scholar
Thomas, M. B. & Willis, A. J. (1998). Biocontrol – risky but necessary? Trends in Ecology & Evolution, 13, 325329.Google Scholar
Thomas, M. B., Wratten, S. D., & Sotherton, N. W. (1991). Creation of “island” habitats in farmland to manipulate populations of beneficial arthropods: Predator densities and emigration. Journal of Applied Ecology, 28, 906917.Google Scholar
Thomas, M. B., Wratten, S. D., & Sotherton, N. W. (1992). Creation of “island” habitats in farmland to manipulate populations of beneficial arthropods: Predator densities and species composition. Journal of Applied Ecology, 29, 524531.Google Scholar
Thomas, P. A. & Room, P. M. (1986). Taxonomy and control of Salvinia molesta. Nature, 320, 581584.Google Scholar
Thompson, W. M. O. (ed.) (2011). The Whitefly, Bemisia tabaci (Homoptera: Aleyrodidae) Interaction with Geminivirus-Infected Host Plants. Dordrecht, Netherlands: Springer.Google Scholar
Tisdell, C. (1990). Economic impact of biological control of weeds and insects. In Critical Issues in Biological Control, ed. Mackauer, M., Ehler, L. E., & Roland, J., pp. 301316. Andover, UK: Intercept.Google Scholar
Tomasetto, F., Tylianakis, J. M., Reale, M., Wratten, S., & Goldson, S. L. (2017). Intensified agriculture favors evolved resistance to biological control. Proceedings of the National Academy of Sciences, U.S.A., 114, 38853890.Google Scholar
Topham, M. & Beardsley, J. W. (1975). Influence of nectar source plants on the New Guinea sugarcane weevil parasite, Lixophaga sphenophori (Villeneuve). Proceedings of the Hawaiian Entomological Society, 22, 145154.Google Scholar
Tothill, J. D., Taylor, T. H. C., & Paine, R. W. (1930). The Coconut Moth in Fiji: A History of its Control by Means of Parasites. London: Imperial Bureau of Entomology.Google Scholar
Tracy, E. F. (2014). The promise of biological control for sustainable agriculture: A stakeholder-based analysis. Journal of Science Policy & Governance, 5 (1).Google Scholar
Traveset, A. (1990). Bruchid egg mortality on Acacia farnesiana caused by ants and abiotic factors. Ecological Entomology, 15, 463467.Google Scholar
Trouvelot, B. (1931). Recherches sur les parasites et predateurs attaquant le doryphore en Amerique du Nord. Annales des Épiphyties, 17, 408445.Google Scholar
Tschumi, M., Albrecht, M., Bärtschi, C., Collatz, J., Entling, M. H., & Jacot, K. (2016). Perennial, species-rich wildflower strips enhance pest control and crop yield. Agriculture, Ecosystems & Environment, 220, 97103.Google Scholar
Tumlinson, J. H., Lewis, W. J., & Vet, L. E. M. (1993). How parasitic wasps find their hosts. Scientific American, 268, 100106.Google Scholar
Turner, C. E. & McEvoy, P. B. (1995). Tansy ragwort (Senecio jacobaea L., Asteraceae). In Biological Control in the Western United States, ed. Nechols, J. R., Andres, L. A., Beardsley, J. W., Goeden, R. D., & Jackson, C. G., pp. 264269. Oakland: University of California, Publication 3361.Google Scholar
Uka, D., Hiraoka, T., & Iwabuchi, K. (2006). Physiological suppression of the larval parasitoid Glyptapanteles pallipes by the polyembryonic parasitoid Copidosoma floridanum. Journal of Insect Physiology, 52, 11371142.Google Scholar
United Nations. (2015). Sustainable development goals: 17 goals to transform our world. Retrieved from www.un.org/sustainabledevelopment/sustainable-development-goals/Google Scholar
Urbaneja, A., Gonzalez-Cabrera, J., Arno, J., & Gabarra, R. (2012). Prospects for the biological control of Tuta absoluta in tomatoes of the Mediterranean basin. Pest Management Science, 68, 12151222.Google Scholar
van Alphen, J. J. M. & Jervis, M. A. (1996). Foraging behaviour. In Insect Natural Enemies: Practical Approaches to their Study and Evaluation, ed. Jervis, M. & Kidd, N., pp. 162. London: Chapman & Hall.Google Scholar
van den Bosch, R. & Hagen, K. S. (1966). Predaceous and parasitic arthropods in California cotton fields. California Agricultural Experiment Station Bulletin, 860.Google Scholar
van den Bosch, R., Messenger, P. S., & Gutierrez, A. P. (1982). An Introduction to Biological Control. New York: Plenum Press.Google Scholar
van den Bosch, R. & Stern, V. (1969). The effect of harvesting practices on insect populations in alfalfa. Proceedings of the Tall Timbers Conference on Ecological Animal Control, 1, 4754.Google Scholar
van den Bosch, R. & Telford, A. D. (1964). Environmental modification and biological control. In Biological Control of Insect Pests and Weeds, ed. DeBach, P. H., pp. 459488. New York: Reinhold Publishing Corporation.Google Scholar
Van de Peer, Y., Ben-Ali, A., & Meyer, A. (2000). Microsporidia: Accumulating molecular evidence that a group of amitochondriate and suspectedly primitive eukaryotes are just curious fungi. Gene, 246, 18.Google Scholar
Van Driesche, R. G. & Bellows, T. S. (eds.) (1993). Steps in Classical Arthropod Biological Control. Lanham, MD: Entomological Society of America.Google Scholar
Van Driesche, R. G., Carruthers, R. I., Center, T., Hoddle, M. S., Hough-Goldstein, J., Morin, L., Smith, L., Wagner, D. L., Blossey, B., Brancatini, V., et al. (2010). Classical biological control for the protection of natural ecosystems. Biological Control, 54, S2S33.Google Scholar
Van Driesche, R. G., Hoddle, M., & Center, T. E. (2008). Control of Pests and Weeds by Natural Enemies: An Introduction to Biological Control. Chichester, UK: Wiley-Blackwell.Google Scholar
Van Driesche, R. G., Pratt, P. D., Center, T. D., Rayamajhi, M. B., Tipping, P. W., Purcell, M., Fowler, S., Causton, C., Hoddle, M. S., Kaufman, L., et al. (2016). Cases of biological control restoring natural systems. In Integrating Biological Control into Conservation Practice, ed. Van Driesche, R., Simberloff, D., Blossey, B., Causton, C., Hoddle, M., Marks, C. O., Heinz, K., Wagner, D., & Warner, K., pp. 208246. Chichester, UK: John Wiley & Sons.Google Scholar
Van Klinken, R. D. & Edwards, O. (2002). Is host specificity of weed biological control agents likely to evolve rapidly following establishment? Ecology Letters, 5, 590596.Google Scholar
van Lenteren, J. C. (1980). Evaluation of control capabilities of natural enemies: Does art have to become science? Netherlands Journal of Zoology, 30, 369381.Google Scholar
van Lenteren, J. C. (2000). Success in biological control of arthropods by augmentation of natural enemies. In Biological Control: Measures of Success, ed. Gurr, G. & Wratten, S., pp. 77103. Dordrecht, Netherlands: Kluwer Academic Publishers.Google Scholar
van Lenteren, J. C. (2003). Need for quality control of mass-produced biological control agents. In Quality Control and Production of Biological Control Agents: Theory and Testing Procedures, ed. van Lenteren, J. C., pp. 118. Wallingford, UK: CABI Publishing.Google Scholar
van Lenteren, J. C. (2012). The state of commercial augmentative biological control: Plenty of natural enemies, but a frustrating lack of uptake. BioControl, 57, 120.Google Scholar
van Lenteren, J. C., Bale, J., Bigler, F., Hokkanen, H. M. T., & Loomans, A. J. M. (2006). Assessing risks of releasing exotic biological control agents of arthropod pests. Annual Review of Entomology, 51, 609634.Google Scholar
van Lenteren, J. C., Bolckmans, K., Köhl, J., Ravensberg, W., & Urbaneja, A. (2018). Biological control using invertebrates and microorganisms: Plenty of new opportunities. BioControl, 63(1), 3959.Google Scholar
van Lenteren, J. C., Hale, A., Klapwijk, J. N., Van Schelt, J., & Steinberg, S. (2003). Guidelines for quality control of commercially produced natural enemies. In Quality Control and Production of Biological Control Agents: Theory and Testing Procedures, ed. van Lenteren, J. C., pp. 265303. Wallingford, UK: CABI Publishing.Google Scholar
van Lenteren, J. C. & Martin, N. A. (1999). Biological control of whiteflies. In Integrated Pest and Disease Management in Greenhouse Crops, ed. Albajes, R., Gullino, M. Lodovica, van Lenteren, J. C., & Elad, Y., pp. 202216. Dordrecht, Netherlands: Kluwer Academic Publishers.Google Scholar
van Lenteren, J. C., Roskam, M. M., & Timmer, R. (1997). Commercial mass production and pricing of organisms for biological control of pests in Europe. Biological Control, 10, 143149.Google Scholar
Vega, F. E. & Kaya, H. K. (2012). Insect Pathology, 2nd edn. San Diego, CA: Academic Press.Google Scholar
Viggiani, G. (1964). La specializzazione entomoparassitica in alcuni Eulofidi (Hym., Chalcidoidea). Entomophaga, 9, 111118.Google Scholar
Vittum, P. J., Villani, M. G., & Tashiro, H. (1999). Turfgrass Insects of the United States and Canada, 2nd edn. Ithaca, NY: Comstock Publishing Associates.Google Scholar
Waage, J. (1990). Ecological theory and the selection of biological control agents. In Critical Issues in Biological Control, ed. Mackauer, M., Ehler, L. E., & Roland, J., pp. 135157. Andover, UK: Intercept.Google Scholar
Waage, J. (1995). The use of exotic organisms as biopesticides: Some issues. In Biological Control, Benefits and Risks, ed. Hokkanen, H. M. T. & Lynch, J. M., pp. 93100. Cambridge: Cambridge University Press.Google Scholar
Wäckers, F. L. & van Rijn, P. C. J. (2012). Pick and mix: Selecting flowering plants to meet the requirements of target biological control insects. In Biodiversity and Insect Pests: Key Issues for Sustainable Management, ed. Gurr, G. M., Wratten, S. D., Snyder, W. E., & Read, D. M. Y., pp. 139165. Chichester, UK: Wiley-Blackwell.Google Scholar
Wagner, D. L., Peacock, J. W., Carther, J. L., & Talley, S. E. (1996). Field assessment of Bacillus thuringiensis on nontarget Lepidoptera. Environmental Entomology, 25, 14441454.Google Scholar
Wang, Z.-Y., He, K.-L., Zhang, F., Lu, X., & Babendreier, D. (2014). Mass rearing and release of Trichogramma for biological control of insect pests of corn in China. Biological Control, 68, 136144.Google Scholar
Wapshere, A. J. (1989). A testing sequence for reducing rejection of potential biological control agents for weeds. Annals of Applied Biology, 114, 515526.Google Scholar
Webster, F. M. (1909). Investigations of Toxoptera graminum and its parasites. Annals of the Entomological Society of America, 2, 6787.Google Scholar
Weller, D. M., Raaijmakers, J. M., McSpadden Gardener, B. B., & Thomashow, L. S. (2002). Microbial populations responsible for specific soil suppressiveness to plant pathogens. Annual Review of Phytopathology, 40, 309348.Google Scholar
Weseloh, R. M. (1998). Possibility for recent origin of the gypsy moth (Lepidoptera: Lymantriidae) fungal pathogen Entomophaga maimaiga (Zygomycetes: Entomophthorales) in North America. Environmental Entomology, 27, 171177.Google Scholar
Wetzstein, H. Y. & Phatak, S. C. (1987). Scanning electron microscopy of the uredinial stage of Puccinia canaliculata on yellow nutsedge, Cyperus esculentus (Cyperaceae). American Journal of Botany, 74, 100106.Google Scholar
Whipps, J. M. & Davies, K. G. (2000). Success in biological control of plant pathogens and nematodes by microorganisms. In Biological Control: Measures of Success, ed. Gurr, G. & Wratten, S., pp. 231269. Dordrecht, Netherlands: Kluwer Academic Publishers.Google Scholar
White, G. (1997). Population ecology and biological control of weeds. In Biological Control of Weeds: Theory and Practical Application, ed. Julien, M. & White, G. (ACIAR Monograph No. 49), pp. 3945. Canberra: Australian Centre for International Agricultural Research.Google Scholar
Wijnands, F. G. & Kroonen-Backbier, B. M. A. (1993). Management of farming systems to reduce pesticide inputs: The integrated approach. In Modern Crop Protection: Developments and Perspectives, ed. Zadoks, J. C., pp. 227234. Wageningen, Netherlands: Wageningen Pers.Google Scholar
Williams, T., Arredondo-Bernal, H. C., & Rodriguez-del-Bosque, L. A. (2013). Biological pest control in Mexico. Annual Review Entomology, 58, 119140.Google Scholar
Williamson, M. & Fitter, A. (1996). The varying success of invaders. Ecology 77, 16611666.Google Scholar
Winkler, K., Wäckers, F., Bukovinszkine-Kissa, F., & van Lenteren, J. (2006). Sugar resources are vital for Diadegma semiclausum fecundity under field conditions. Basic and Applied Ecology, 7, 133140.Google Scholar
Winston, R. L., Schwarzländer, M., Hinz, H. L., Day, M. D., Cock, M. J. W., & Julien, M. H. (eds.) (2014). Biological Control of Weeds: A World Catalogue of Agents and Their Target Weeds, 5th edn. Morgantown, WV: US Department of Agriculture, Forest Service, Forest Health Technology Enterprise Team.Google Scholar
Wood, A. R. (2012). Uromycladium tepperianum (a gall-forming rust fungus) causes a sustained epidemic on the weed Acacia saligna in South Africa. Australasian Plant Pathology, 41, 255261.Google Scholar
Yandoc-Ables, C. B., Rosskopf, E. N., & Charudattan, R. (2006). Plant pathogens at work: Progress and possibilities for weed biocontrol. Part 2: Improving weed control efficiency. Retrieved from www.apsnet.org/publications/apsnetfeatures/Pages/WeedBiocontrolPart2.aspx.Google Scholar
Yandoc-Ables, C. B., Rosskopf, E. N., & Charudattan, R. (2007). Plant pathogens at work: Progress and possibilities for weed biocontrol: classical versus bioherbicidal approaches. Plant Health Progress. DOI: 10.1094/PHP-2007-0822-01-RV.Google Scholar
Yaninek, J. S. & Hanna, R. (2003). Cassava green mite in Africa – A unique example of successful classical biological control of a mite pest on a continental scale. In Biological Control in IPM Systems in Africa, ed. Neuenschwander, P., Langewald, J. & Borgemeister, C., pp. 6176. Wallingford, UK: CABI Publishing.Google Scholar
Zelazny, B., Lolong, A., & Pattang, B. (1992). Oryctes rhinoceros (Coleoptera, Scarabaeidae) populations suppressed by a baculovirus. Journal of Invertebrate Pathology, 59, 6168.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • Bibliography
  • Ann E. Hajek, Cornell University, New York, Jørgen Eilenberg, University of Copenhagen
  • Book: Natural Enemies
  • Online publication: 06 July 2018
  • Chapter DOI: https://doi.org/10.1017/9781107280267.023
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.

  • Bibliography
  • Ann E. Hajek, Cornell University, New York, Jørgen Eilenberg, University of Copenhagen
  • Book: Natural Enemies
  • Online publication: 06 July 2018
  • Chapter DOI: https://doi.org/10.1017/9781107280267.023
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.

  • Bibliography
  • Ann E. Hajek, Cornell University, New York, Jørgen Eilenberg, University of Copenhagen
  • Book: Natural Enemies
  • Online publication: 06 July 2018
  • Chapter DOI: https://doi.org/10.1017/9781107280267.023
Available formats
×